You are on page 1of 23

Contrib Mineral Petrol (2017) 172:97

DOI 10.1007/s00410-017-1418-1

ORIGINAL PAPER

Lower crustal hydrothermal circulation at slow‑spreading ridges:


evidence from chlorine in Arctic and South Atlantic basalt glasses
and melt inclusions
Froukje M. van der Zwan1,6   · Colin W. Devey1   · Thor H. Hansteen1 ·
Renat R. Almeev2 · Nico Augustin1 · Matthias Frische1 · Karsten M. Haase3 ·
Ali Basaham4 · Jonathan E. Snow5 

Received: 29 March 2017 / Accepted: 25 October 2017


© Springer-Verlag GmbH Germany 2017

Abstract  Hydrothermal circulation at slow-spreading 50–430 ppm and ca. 40–700 ppm for the SMAR and Gakkel
ridges is important for cooling the newly formed lithosphere, Ridge, respectively, whereas SMAR melt inclusions contain
but the depth to which it occurs is uncertain. Magmas which between 20 and 460 ppm Cl. Compared to elements of simi-
stagnate and partially crystallize during their rise from the lar mantle incompatibility (e.g. K, Nb), Cl-excess (Cl/Nb or
mantle provide a means to constrain the depth of circula- Cl/K higher than normal mantle values) of up to 250 ppm in
tion because assimilation of hydrothermal fluids or hydro- glasses and melt inclusions are found in 75% of the samples
thermally altered country rock will raise their chlorine (Cl) from both ridges. Cl-excess is interpreted to indicate assimi-
contents. Here we present Cl concentrations in combina- lation of hydrothermal brines (as opposed to bulk altered
tion with chemical thermobarometry data on glassy basaltic rock or seawater) based on the large range of Cl/K ratios in
rocks and melt inclusions from the Southern Mid-Atlantic samples showing a limited spread in H ­ 2O contents. Resorp-
Ridge (SMAR; ~ 3 cm year−1 full spreading rate) and the tion and disequilibrium textures of olivine, plagioclase and
Gakkel Ridge (max. 1.5 cm year−1 full spreading rate) in clinopyroxene phenocrysts and an abundance of xenocrysts
order to define the depth and extent of chlorine contami- and gabbroic fragments in the SMAR lavas suggest multi-
nation. Basaltic glasses show Cl-contents ranging from ca. ple generations of crystallization and assimilation of hydro-
thermally altered rocks that contain these brines. Calculated
pressures of last equilibration based on the major element
Communicated by Jochen Hoefs. compositions of melts cannot provide reliable estimates of
the depths at which this crystallization/assimilation occurred
Electronic supplementary material  The online version of
this article (https://doi.org/10.1007/s00410-017-1418-1) contains
as the assimilation negates the assumption of crystallization
supplementary material, which is available to authorized users. under equilibrium conditions implicit in such calculations.
Clinopyroxene–melt thermobarometry on rare clinopyroxene
* Froukje M. van der Zwan phenocrysts present in the SMAR magmas yield lower crus-
fzwan@geomar.de
tal crystallization/assimilation depths (10–13 km in the seg-
1
Geomar Helmholtz Centre for Ocean Research Kiel, Kiel, ment containing clinopyroxene). The Cl-excesses in SMAR
Germany melt inclusions indicate that assimilation occurred before
2
Leibniz Universität Hannover, Institute of Mineralogy, crystallization, while also homogeneous Cl in melts from
Hannover, Germany Gakkel Ridge indicate Cl addition during magma chamber
3
GeoZentrum Nordbayern, Universität Erlangen-Nürnberg, processes. Combined, these observations imply that hydro-
Erlangen, Germany thermal circulation reaches the lower crust at slow-spreading
4
Faculty of Marine Science, King Abdulaziz University, ridges, and thereby promotes cooling of the lower crust. The
Jeddah, Saudi Arabia generally lower Cl-excess at slow-spreading ridges (com-
5
Department of Earth and Atmospheric Sciences, University pared to fast-spreading ridges) is probably related to them
of Houston, Houston, USA having few if any permanent magma chambers. Magmas
6
Present Address: Institute für Geowissenschaften, therefore do not fractionate as extensively in the crust, pro-
Christian-Albrechts-Universität Kiel, Kiel, Germany viding less heat for assimilation (on average, slow-spreading

13
Vol.:(0123456789)
97   Page 2 of 23 Contrib Mineral Petrol (2017) 172:97

ridge magmas have higher Mg#), and hydrothermal systems (e.g. < 1600 m in Alt and Bach 2006) or from tectonically
are ephemeral, leading to lower total degrees of crustal alter- exposed parts on the lower crust (Gillis et al. 1993), which
ation and more variation in the amount of Cl contamina- have thus been exposed to ambient seawater for extended
tion. Hydrothermal plumes and vent fields have samples in periods of time.
close vicinity that display Cl-excess, mostly of > 25 ppm, Deep hydrothermal circulation could potentially be traced
which thus can aid as a guide for the exploration of (active by detecting it effects on the chlorine budget of erupted mag-
or extinct) hydrothermal vent fields on the axis. mas if those magmas assimilated the hydrothermal fluids
themselves or hydrothermally altered country rock at depth
Keywords  Hydrothermal circulation · (Ultra)slow- (Michael and Schilling 1989; Michael and Cornell 1998;
spreading ridges · Crystallization depths · Crustal Kent et al. 1999; Coogan et al. 2003; Gillis et al. 2003; Sun
assimilation · MORB · Chlorine et al. 2007; France et al. 2009; Kendrick et al. 2013; van der
Zwan et al. 2015). Assimilation of surrounding rocks by ris-
ing MORB magmas has been shown to take place in several
Introduction gabbroic complexes (e.g. Bédard et al. 2000 and references
therein; Godard et al. 2009; Drouin et al. 2009; Sanfilippo
When new oceanic lithosphere is formed at a spreading et al. 2014; Fischer et al. 2016), but the depth and/or hydro-
centre, heat is transferred upwards from the mantle either thermal alteration state of the assimilated rocks is uncertain
by advection (magmatism) or tectonism. Although there is so far.
fairly convincing evidence that hydrothermal circulation of Although work on fast-spreading ridges indicates addition
seawater cools the upper parts of this system (e.g. Stein and of Cl relative to trace elements of similar incompatibility
Stein 1994), up to present no consensus on the maximum (e.g. K or Nb) in a reaction zone between the magma and
depth of hydrothermal circulation in newly formed oceanic the overlying hydrothermal circulation cell at the roof of a
crust (e.g. Mottl 2003) or what role hydrothermal circulation magma chamber (Coogan et al. 2003; Gillis et al. 2003; le
plays in the deep cooling of ridges has been reached. Particu- Roux et al. 2006; France et al. 2009; Kendrick et al. 2013),
larly at (ultra)slow-spreading (< 1.2 cm year−1 full spreading previous studies on slow-spreading ridges (e.g. Michael
rate) and slow-spreading ridges (< 1.2 to < 5.5 cm year−1) and Cornell 1998) had not found evidence of Cl addi-
cooling of the lower lithosphere by hydrothermal circulation tion to their magmas. Michael and Cornell (1998) argued
may be relevant, as seafloor is partially produced by tectonic that because fast-spreading ridges generally have shallow
processes (e.g. Dick et al. 2003; Cannat et al. 2006) and the (100–300 MPa ≈ 3–10 km) axial magma chambers whereas
lack of permanent magma chambers and low crustal tem- magmas from a typical slow-spreading ridge (e.g. not anom-
peratures (e.g. Harper 1985; Detrick et al. 1990) may allow alously thick) generally yield high calculated crystallization
faulting to be particularly extensive and to penetrate deeper, pressures (> 300 MPa = > 10 km), hydrothermal circulation
down to the lower crust, potentially providing pathways for (at least at slow-spreading ridges) must occur at pressures
hydrothermal fluids (Harper 1985; Cannat et al. 1991; Mével less than 300 MPa. More recent studies have now found
and Cannat 1991). Cl addition in basalts of some slow-spreading ridges (e.g.
Direct evidence for deep circulation comes from studies the Red Sea Rift) independent of their calculated crystal-
of oceanic drill sections (from crust formed at fast-spreading lization pressures (Jenner and O’Neill 2012; van der Zwan
ridges) and from ophiolites (e.g. Gregory and Taylor 1981; et al. 2015) and advances in analytical methods mean that
Stakes and Vanko 1986; Nehlig and Juteau 1988; Lecuyer the inherently lower Cl contents of the relatively less frac-
and Reynard 1996; Kawahata et al. 2001; Coogan 2003; tionated slow-spreading ridge magmas have become less of
Nicolas et al. 2003), both of which display high-temperature an analytical challenge.
alteration of the lower crust. Deep hydrothermal circulation To investigate how deep hydrothermal circulation reaches
at fast-spreading ridges is further indicated by thermody- at slower-spreading ridges, we performed high-precision Cl
namic modelling (Cherkaoui et al. 2003; Maclennan et al. measurements (van der Zwan et al. 2012) on basaltic glasses
2005; Hasenclever et al. 2014) and the seismic structure of and melt inclusions and combined this with barometric data
the lower crust (Dunn et al. 2000). Evidence from slow- based on glass compositions and on glass and clinopyroxene
spreading ridges include P-wave tomography underneath phenocryst compositions (cf. Ariskin and Barmina 2004;
the TAG hydrothermal field, which indicates hydrother- Putirka 2008). As the composition and structure of the new
mal circulation, related to a detachment fault, extending at lithosphere at slower-spreading ridges is very heterogene-
least 2 km, but potentially 3.5 km deep (Zhao et al. 2012). ous, we selected two end-member ridges for our study: the
Other direct information on the depth of penetration of magmatically robust southern Mid-Atlantic Ridge (SMAR)
hydrothermal circulation at slow-spreading ridges is only at 7–10°S (a slow-spreading ridge with ~  3  cm  year−1)
available from drill-cores that have a limited depth extent previously interpreted to show varying mantle sources

13
Contrib Mineral Petrol (2017) 172:97 Page 3 of 23  97

and crystallization pressures (0–800 MPa; Almeev et al. and A2 as these samples display the largest range in calcu-
2008; Hoernle et al. 2011) and the Arctic Gakkel Ridge lated last crystallization pressures (200–900 MPa; Almeev
at 6°W–85°E (an ultraslow-spreading ridge with max. et  al. 2008). Segment A1 has a typical slow-spreading
1.5 cm year−1) which has a strongly tectonic character and ridge character with a well-developed deep rift valley of
a very thin crust (Coakley and Cochran 1998; Dick et al. > 2950 m, while segment A2 shows a shallower axial high
2003; Michael et al. 2003). Note that at slow- and espe- and is 2100–3000 m deep (Fig. 1a; Bruguier et al. 2003).
cially ultraslow spreading ridges classical “Penrose-type” Segment A2 is associated with a thick seismically deter-
magmatic crust (e.g. Vine and Moores 1972) may not make mined crust of ~ 11 km, compared to ~ 5 km under A1
up the whole newly formed lithosphere (e.g. Cannat 1996) (Minshull et al. 1998; Bruguier et al. 2003). Segment A1
and may even be totally absent (e.g. Sauter et al. 2013). In displays microseismicity at 2–4 km depth and teleseismic-
this paper we therefore use the term “crust” to refer to the ity (Grevemeyer et al. 2013), while segment A2 shows little
wholly magmatic upper part of the lithosphere, if present. earthquake activity, thought to indicate a higher tempera-
The term “lithosphere” refers to the combination of crustal ture gradient and therefore a shallower brittle/ductile transi-
and/or mantle rocks which make up the new plate. tion in the crust (Fig. 1a; Devey et al. 2010). Presently, A2
appears to be in a period of reduced magmatic activity: the
youngest lava flows are tectonized and yield ages around
Geological background 1000 years and no signs of active hydrothermalism have
been found (Devey et al. 2010; Haase et al. 2016). The sam-
Southern Mid‑Atlantic Ridge ples used here are well studied in terms of their major and
trace element and isotope compositions and crystallization
The southern Mid-Atlantic Ridge (SMAR) between 7° and pressures (Möller 2002; Almeev et al. 2008; Paulick et al.
10°S, east of Ascension island, is a mature ridge with slow 2010; Hoernle et al. 2011). The geochemical studies show
spreading rates of 31–32  mm  year −1 full spreading rate mixing between a depleted unradiogenic MORB source
(DeMets et al. 2010; Fig. 1a). The area is bounded to the and radiogenic, trace element enriched HIMU-type sources
north by the Ascension Fracture Zone and to the south by (Hoernle et al. 2011). The enriched samples display higher
the Bode Verde Fracture Zone. The region is divided into ­H2O contents and nearly isobaric calculated crystallization
four segments (A1–A4; Bruguier et al. 2003) that are sepa- pressures (100–300 MPa) while the more depleted samples
rated by overlapping spreading centres and show varying show lower H ­ 2O contents and a range of calculated last crys-
morphologies. We chose to study samples from segments A1 tallization pressures (200–900 MPa; Almeev et al. 2008).

Fig. 1  Overview of studied
areas on the a Southern Mid-
Atlantic Ridge (SMAR) and b
Gakkel Ridge. The SMAR study
area is subdivided in various
segments (after Bruguier et al.
2003; samples are from A1 and
A2). The Gakkel Ridge is seg-
mented into three zones (after
Michael et al. 2003): western
volcanic zone (WVZ), sparsely
magmatic zone (SMZ) and east-
ern volcanic zone (EVZ). White
dots indicate earthquake hypo-
centers with M ≥ 3 (from ISC,
1960 to present). Black stars
indicate hydrothermal activity
after German et al. (2002) and
Baker et al. (2004). Spreading
rates (after DeMets et al. 2010)
are indicated by arrows

13
97   Page 4 of 23 Contrib Mineral Petrol (2017) 172:97

Hydrothermal circulation in the studied area is known from per rock under a binocular microscope based on their fresh
the Nibelungen hydrothermal field between A1 and A2 and appearance and lack of visible vesicles, alteration rims or
from two indications of plume activity at 7°57′S and 8°17′S minerals. The mounted glass was prepared, polished and
on A1 (Fig. 1a; Devey et al. 2005; Melchert et al. 2008). then cleaned with Milli-Q water following the procedures in
van der Zwan et al. (2012) to avoid Cl contamination. Opti-
Gakkel Ridge cally clear SMAR Ol, Pl and Cpx minerals were prepared
following the same procedure. For selected SMAR samples
The Gakkel Ridge (6°W–85°E) shows the lowest orthogonal with elevated Cl/Nb glass contents, Ol and Pl minerals with
spreading rate on Earth at 11–13 mm year−1 full spread- melt inclusions were mounted in UHU Hart glue mixed with
ing rate (DeMets et al. 2010) and lacks transform offsets acetone and doubly polished sections (~ 60 µm thick) were
(Fig. 1b). The Gakkel Ridge has a very deep rift valley prepared. Melt inclusions were selected based on homoge-
(> 4600 m) and a thin seismically determined crustal thick- neity and large sizes/thicknesses and with a good surface.
ness (< 4 km; Coakley and Cochran 1998; Jokat et al. 2003). Major elements and Cl measurements were carried out
The Gakkel Ridge can be subdivided into three different with a Jeol JXA-8200 ‘‘Superprobe’’ electron microprobe
magmato-tectonic units (Fig. 1b; Michael et al. 2003). The at GEOMAR, Kiel, using an acceleration voltage of 15 kV.
Western Volcanic Zone (WVZ; 7°W–3°E) is a volcanic Chlorine and potassium (K) were measured with a beam
area with many small cones on the ridge and displays a current of 80 nA and a beam diameter of 10 µm using the
larger extent of melting than the other two segments. In the mapping technique described in van der Zwan et al. (2012,
Sparsely Magmatic Zone (SMZ; 3°E–28°E) in contrast, the 2015). The Cl and K results are the average of 2–8 meas-
seafloor consists mostly of peridotite with only rare basalts urements per sample and have an uncertainty of < 3 and
or volcanic features. The Eastern Volcanic Zone (EVZ; < 20 ppm for Cl and K, respectively (2SE). For the Cl and
29°E–85°E) contains large (~ 30 km) widely spaced vol- K measurements of the melt inclusions, the method was
canic centres of which some show recent activity. Basalts slightly adapted by a reduced beam current of 50 nA and a
from the Gakkel Ridge are generally enriched in trace ele- mapping repetition of 14 times, in order to avoid a strong Cl
ments but with a large variability between the different signal of the mounting material through the thin samples.
segments and also between samples (Michael et al. 2003). Internal glass standards (138-3, 147-2, 157-3; van der Zwan
Isotopically, two distinct sources were identified: samples et al. 2012) show a reproducibility of on average 3.2 ppm
from the WVZ have a DUPAL-signature attributed to a (2SD) for this method. Nevertheless, an additional signal
contribution from the sub-continental lithospheric mantle, from the mounting material is visible in the Cl data of the
while samples from the EVZ have a signature similar to melt inclusions and corrected for, but results in a higher
Atlantic–Pacific MORBs (Goldstein et al. 2008) with the uncertainty of 19.2 ppm 2SD. This higher uncertainty does,
boundary between these mantle domains lying in the SMZ. however, not influence the trends observed in the melt inclu-
Hydrothermal activity at Gakkel Ridge seems to be high sions, neither the main conclusions derived from Cl data of
compared to other (ultra)slow-spreading ridges (Edmonds the melt inclusions. The K measurements are not affected
et al. 2003; Michael et al. 2003). Although only two active by the mounting material. Any measurements that show dis-
vent sites were confirmed by visual observations (Aurora, crepancies (e.g. inhomogeneous Cl, mostly due to bubble
85°E), 7–11 more vent sites were inferred by light scattering, forming) were discarded.
temperature and Mn concentration data in the water column, Major elements of the glasses were measured with a defo-
from which we here consider the seven most certain ones cused spot of 5 µm and a beam current of 10 nA. Counting
(Fig. 1b; Edmonds et al. 2003; Baker et al. 2004; Pontbriand times were 20/10 s (peak/background left and right of the
et al. 2012). Teleseismic earthquakes are present along the peak each 10 s) for all elements apart for 30/15 s for MnO.
whole ridge (Fig. 1b). ­Na2O was measured first to avoid loss by heating. For cali-
bration and monitoring of data quality, we used natural ref-
erence samples from the Smithsonian Institute (Jarosewich
Analytical methods et al. 1980). The average was taken of nine analyses per
sample to assure homogeneity. Relative analytical preci-
Chlorine, major, trace, H­ 2O and C­ O2 measurements and sion is generally < 2.5%, but up to 5% for N
­ a2O and ~ 30%
pressure calculations were performed on glassy basaltic for MnO and ­P2O5. Additionally, available (literature) data
samples from collections of expeditions R/V Meteor 41/2 for the samples were used; for the SMAR, data are utilized
(SMAR; e.g. Möller 2002; Almeev et al. 2008; Hoernle et al. from Möller (2002) and data measured with a CAMECA
2011) and AMORE 2001 (Gakkel Ridge; e.g. Michael et al. SX-50 at GEOMAR and a JEOL JXA8900 Superprobe at the
2003; Goldstein et al. 2008). For each basaltic glass meas- Institut für Geowissenschaften, University of Kiel, partially
urement, three pieces of glass > 2 mm across were selected presented in Haase et al. (2016). Settings were the same as

13
Contrib Mineral Petrol (2017) 172:97 Page 5 of 23  97

above and accuracy and precision are better than 5% except for basaltic compositions: ρ = 2819 − 20.8·H2O concen-
for MnO and ­P2O3. Additional major element data measured tration. The ­H2O and ­CO2 concentrations were calculated
on a Jeol JXA 8900RL electron probe microanalyzer at the based on the Lambert–Beer law using the peak height, which
University of Mainz were used for Gakkel Ridge basalts. was determined by reference to a straight tangential base
Mineral point analyses and transects were measured with a line. The average values of three measurements per sample
focused beam spot. Plagioclase and Cpx were measured with were used to calculate the H ­ 2O and C
­ O2 contents of the
a beam current of 20 nA and counting times of 20/10 s, apart glasses, with a standard deviation usually less than 0.02 and
for ­K2O in Pl (40/20 s) and ­Cr2O3 and ­TiO2 in Cpx (30/15 s). 0.003 wt%, respectively.
Olivine was measured with a beam current of 100 nA and Pressure and temperature conditions at the point of last
20/10 s counting times for S ­ iO2, ­Al2O3, MgO, FeO, 30/15 s crystallization equilibrium were calculated by two differ-
for MnO, ­Cr2O3, 60/30 s for CaO and 40/20 s for NiO. ent methods. Pressures for all samples were calculated with
For consistency, trace elements in all glass samples were the software COMAGMAT version 3.57 that is based on
analysed by LA-ICP-MS at GEOMAR, Kiel, using a 193 nm empirical models calibrated with an experimental data-
Excimer laser system (GeoLasPro, Coherent) coupled with base (Ariskin and Barmina 2004). Here we used inversed
a double-focusing, high-resolution magnetic sector mass modelling (see Almeev et al. 2008) of the major elements
spectrometer (AttoM, Nu-Instruments) under hot plasma and ­H2O concentrations of the glasses. The effects of small
conditions [NAI > 10; ThO/Th ~ 0.03%; details in Fietzke amounts of ­H2O on crystallization were taken into account
and Frische (2016)]. Spot analyses of samples were done by incorporating recent experimental calibration coeffi-
by 30 s ablation at a laser repetition rate of 10 Hz using cients (Almeev et al. 2007a, b, 2012). The pressure models
spot diameter of 90 µm and a fluence of 5 J cm−2. 50 s of of COMAGMAT have an uncertainty of ± 100 MPa and are
gas background data were collected prior to each ablation. comparable with models used in Michael and Cornell (1998)
NISTSRM610 [30 s, 32 µm, 10 Hz, 5 J cm−2; Wise and (Almeev et al. 2008). For the samples that contain Cpx, pres-
Watters 2012] was used for calibration. MPI-DING refer- sures were also calculated using equilibrium exchange mod-
ence samples [ML3B-G and KL2-G; 30 s, 32–90 µm, 10 Hz, els between Cpx and the melt (Putirka et al. 1996; Putirka
5 J cm−2; Jochum et al. (2006)] were used to evaluate the 2008), determined by analysing the average compositions
accuracy during the measurements. Relative analytical pre- of Cpx-rims and their host glasses. Mid-ocean ridge basalts
cision based on reproducibility (2SD) of the standard ref- have typically an ­Fe3+ content of 10–12% (Bézos and Hum-
erence material used for calibration was about 3.5% (see ler 2005; Berry et al. 2017); for the calculations here we
Online Resource 1). Data evaluation has been performed applied a ­Fe3+ content of 12% (Bézos and Humler 2005),
applying the linear regression slope method (Fietzke et al. but variations within these values do not affect the pres-
2008). Ca and Si were used for internal standardization uti- sure calculations. The standard error of estimate (SEE) of
lizing data from EMP analyses. the accuracy of the Cpx–melt model is ± 150 MPa (for an
Glass ­H2O and ­CO2 concentrations were determined extended evaluation of involved errors see Putirka 2008).
using a Fourier transformation infrared (FTIR) Bruker IFS88
spectrometer coupled with an IR Scope II microscope at
the Institute of Mineralogy, Leibniz University of Hanno- Results
ver. ­H2O and C­ O2 concentrations were measured on doubly
polished glass chips (~ 1 mm2 in size) that had a thickness Mineral textures and chemistry
of 150–50 μm, measured with a digital micrometer Mitutoyo
[precision ± 2 μm (Behrens et al. 2009)] by three points Phenocrysts in the SMAR lavas are up to 2 mm in size and
for every polished glass fragment. Absorption spectra in mostly comprise olivine (Ol) and plagioclase (Pl), with
the mid-infrared (MIR) range were collected using a spot minor clinopyroxene (Cpx) observed in only five of the
size of 100 × 100 μm (average of 50 scans) and using the samples. In samples that cover all different chemical com-
following operating conditions: globar light source, KBr positions Ol and Pl show a range of textures from euhedral
beam splitter, MCT (HgCdTe) detector, 4 cm−1 spectral crystals to anhedral grains displaying resorption at the min-
resolution, spectral range 13,000–0 cm−1. The H ­ 2O concen- eral rims and occasionally sieve textures (Fig. 2a–e). A few
tration was measured at the peak that is attributed to the of the Ol grains display undulose extinction (Fig. 2f). Melt
OH stretch vibration (3550 cm−1) using a molar absorption inclusions are present in both Ol and Pl and can be up to
coefficient of 67 L cm−1 mol−1 (Stolper 1982). The C ­ O2 con- ~ 200 µm in diameter; in the Pl they are frequently located
centration was measured at the peak doublet with maxima at along growth zones, but also along healed cracks. Olivine
1430 and 1510 cm−1 using molar absorption coefficient of and Pl crystals often show clustering and intergrowth of
317 L cm−1 mol−1 (Shishkina et al. 2010). Glass densities multiple minerals that probably represent cumulate frag-
were calculated using the equation of Yamashita et al. (1997) ments (Fig. 2c–h).

13
97   Page 6 of 23 Contrib Mineral Petrol (2017) 172:97

Fig. 2  Photomicrographs
of thin sections from SMAR
samples 149-1 (a, c–f, h), 142-7
(b) and 145-3 (g) in cross-
polarized light (a–d, f–h) and
plane-polarized light (e). a, b,
h Euhedral Ol occurs together
with highly dissolved and skel-
etal Pl and/or Ol. c–h Clusters
of multiple Pl (c), or Pl and Ol
(d–h) with complex intergrowth
structure of Pl and Ol (e, g) and
undulose extinction in Ol (f)
indicate xenocrystic fragments
that formed by in situ crustal
growth of minerals together and
subsequent assimilation of these
fragments

13
Contrib Mineral Petrol (2017) 172:97 Page 7 of 23  97

Fig. 3  Mineral composition data. a Transects over Ol grains from ▸


rim to core show unzoned Ol together with normally and slightly
reversely zoned Ol. b Fo-content of the Ol plotted against the Fo-con-
tent of Ol that would be in equilibrium with their host glasses (apply-
ing the equilibrium constant of 0.3  ±  0.03 from Roeder and Emslie
1970). Olivines in equilibrium fall onto the one-to-one line or within
the red field that represents the variation of the equilibrium constant.
c Transects over Cpx crystals show various zonation patterns from
limited zoned to complexly zoned crystals. d Plagioclase transects
display scatter showing oscillatory zoning, but display a generally
reversely zoned trend (147-2) and unzoned trend (143-2)

The forsterite (Fo) contents in Ol vary between 83.4 and


91.8%, with intra-sample variations up to 5.5% both between
different grains and within single Ol grains. Chemical pro-
files from core to rim over several selected Ol reveal (1)
crystals with a constant composition, (2) normally zoned
crystals, of which some may include diffusion profiles
(e.g. in 143-2 and 149-1) and (3) crystals with a slightly
reverse zonation (e.g. in 143-2, 147-2; Fig. 3a). Several Ol
rim compositions are too Fo-rich to be in equilibrium with
their host melts (Fig. 3b; using the equilibrium constant of
0.30 ± 0.03 from Roeder and Emslie (1970) and 12% F ­ e3+ in
the glasses (Bézos and Humler 2005) points to equilibrium
at Fo-contents of 79.4–89.6%) implying a xenocrystic origin
for these Ol. Also here applying a slightly different ­Fe3+
content does not affect this observation. Olivines in disequi-
librium are found for both incompatible element enriched
and depleted samples. The rare Cpx that occur in samples
147-2, 149-1, 149-4, 162-1 and 162-2 have a composition
of ­Wo36–41En51–55Fs6–10 and rim to core profiles display
zoning in some cases (Fig. 3c). Two profiles over Pl show
oscillatory zoning around a constant composition or slight
reversed-zoning (Fig. 3d).

Composition of the glasses

The major and trace elements together with the Cl data


are given in Online Resources 2 (SMAR) and 3 (Gak-
kel Ridge). The Cl concentration values range from 43
to 704 ppm for Gakkel Ridge and 52–429 ppm for the
SMAR (± < 3 ppm 2S.E.), although most values lie below
200 ppm (Fig. 4a). This range is similar to literature data
for other slow-spreading ridges (Michael and Cornell
1998; PetDB, Lehnert et  al. 2000; Jenner and O’Neill
2012). In the majority of both SMAR and Gakkel sam-
ples there is a negative correlation between Cl and MgO
(Fig. 4a). This trend is, however, poor (R2 = 0.22; n = 117)
and not parallel to a single fractionation model [(log(Cl8/
Cl) = 0.12·(MgO-8) from Weaver and Langmuir (1990),
assuming unfractionated Cl concentrations of 50, 100 or
150 ppm at 8 wt% MgO; Fig. 4a]. Several Cl-rich samples correlations are seen between Cl and several incompatible
from both SMAR and Gakkel Ridge lie well above the elements, e.g. K, Rb, Ba, Th, U, Nb, Ta, La, Ce, Pr, Nd
values that could be explained by fractionation. Positive and ­H 2O, with the best correlation existing between Cl

13
97   Page 8 of 23 Contrib Mineral Petrol (2017) 172:97

Fig. 4  Chlorine composi-
tions of SMAR and Gakkel
Ridge basalt glasses compared
to major and trace element
compositions. The analytical
error is smaller than the symbol
sizes. a Cl vs. MgO compared
to fractionation trends for Cl
compositions of 50, 100 or
150 ppm at 8 wt% MgO (black
lines). Cl vs. Nb for b Gakkel
Ridge and c SMAR samples
together with lines of various
Cl/Nb ratios and compared to
samples from the Red Sea (van
der Zwan et al. 2015). The local
mantle Cl/Nb ratio (see “Dis-
cussion”) for the two locations
is highlighted. Note that many
samples lie above these mantle
values. d Chlorine concentra-
tions are related to radiogenic
isotope ratios (i.e. 87Sr/86Sr;
Hoernle et al. 2011). e ­H2O and
Ce contents are relatively well
correlated. f Cl/K and Cl/Nb are
correlated for each sample suite,
but display different trends
between the suites. g Variations
in Nb/K between the samples
suites are corresponding to Ce/
Pb

and Nb (R2 = 0.87; Fig. 4b, c), the element whose mantle have not lost or gained significant amounts of water since
incompatibility is most similar to Cl (Sun et al. 2007). leaving the melting zone (cf. Michael 1995).
Samples from the SMAR also show a correlation between To study Cl addition to the magmas, we examined the
Cl and radiogenic Sr, Nd and Pb isotopes (isotope data ratios of Cl to elements of similar incompatibility; wher-
from Hoernle et  al. (2011); Fig.  4d; Table  1). No cor- ever possible we used Nb (Cl/Nb) as Nb is slightly closer
relation (R2 < 0.2) is observed between Cl and indicators in incompatibility to Cl than K (Sun et al. 2007) and is
of degassing (e.g. C
­ O2/Nb, ­CO2/Ba, cf. Saal et al. 2002; more resistant to the effects of weathering and alteration
Michael and Graham 2015), melting degrees (e.g. ­Na8, cf. processes; for analytical reasons Cl/K is used for the melt
Klein and Langmuir 1987) or continental input (e.g. Ce/ inclusions, and to compare our data to older studies, where
Pb, U/Nb, cf. Hofmann et al. 1986). A good correlation mostly only major elements were measured. Cl/K and Cl/
between ­H2O and Ce (Fig. 4e) suggests that the magmas Nb are positively correlated for each study area although

13
Contrib Mineral Petrol (2017) 172:97 Page 9 of 23  97

Table 1  Pressure calculations Sample Cl/Nb Pressure Putirka et al. KD (Fe–Mg) Pressure COMAG- ∂ Pressure
based on the formulations for (1996) (MPa) MAT (MPa)
Cpx–melt equilibria compared
to COMAGMAT pressure 147-2_1 47.9 361 0.25 670 − 309
calculations
147-2_2 47.9 316 0.24 670 − 354
147-2_3 47.9 324 0.23 670 − 346
149-1_1a 32.8 − 13 0.12 5 − 18
149-1_2a 32.8 53 0.12 5 48
149-4 33.3 375 0.17 5 370
162-1_1 18.3 387 0.20 80 307
162-1_2 18.3 403 0.20 80 323
162-1_3 18.3 373 0.20 80 293
162-2_1 18.1 340 0.24 10 330
162-2_2 18.1 364 0.25 10 354
a
 The low/negative pressures together with the low KD (Fe–Mg) indicate disequilibrium and geologically
non-significant pressures for this sample

each area lies on a different subparallel trend (Fig.  4f). in the host glasses (0.12–0.19). For individual samples,
These trends are consistent with source-derived variations Cl, K and Cl/K of the melt inclusions vary to both higher
in Nb/K between the different sample suites as they also and lower values than their host glass (stars in Fig. 6b). An
show a grouping in their Ce/Pb and Nb/U contents (Fig. 4g). exception are the trace element enriched samples from sta-
SMAR samples reach a maximum Cl/K of 0.3 and Cl/Nb tion 149, which display a larger variation of the melt inclu-
of 64, higher than at most other slow-spreading ridges, but sions in Cl/K, but both Cl and K are considerably lower than
lower than has been observed at fast-spreading ridges or in the host glass and overlap with the more depleted samples
the Red Sea (Fig. 5). The Gakkel Ridge samples, in contrast, (Fig. 6c). The highest Cl/K values in the melt inclusions are
all lie within the previously defined slow-spreading ridge observed in 143-1 and 143-2 that also have the highest Cl/K
field (Michael and Cornell 1998; Fig. 5). High Cl/K ratios of their host glasses (0.19).
at SMAR mostly occur in trace element depleted samples
(e.g. samples with low K/Ti, Fig. 5a), consistent with the Pressures of equilibrium crystallization
negative relation observed between Cl/Nb and incompat-
ible trace elements (e.g. Nb, K, Rb, REE; Fig. 5c). Samples Previous work (Almeev et al. 2008) had used the experi-
from Gakkel Ridge show a range of Cl/Nb values across mentally calibrated COMAGMAT program to determine the
the whole range of trace element concentrations (Fig. 5b). pressures at which the SMAR magmas were last in equilib-
High Cl/Nb is observed in both segments of the SMAR and rium with the three-phase assemblage Ol–Pl–Cpx. Applying
in the WVZ and SMZ of the Gakkel Ridge. The EVZ of the COMAGMAT with improved experimental calibration coef-
Gakkel Ridge in contrast, displays comparatively few high ficients for the effect of water (see method section) to our
Cl/Nb samples. samples yields pressures of last equilibrium crystallization
between 5 and 700 MPa for the SMAR and between 100 and
High‑precision chlorine measurements on SMAR melt 840 MPa for the Gakkel Ridge (Online Resources 2, 3). The
inclusions average pressures are 357 and 426 MPa, respectively, similar
to values found by Michael and Cornell (1998) for slow-
High-precision Cl and K concentrations were measured in spreading ridges which did not display high Cl/K. SMAR
melt inclusions in both Ol and Pl crystals in selected high Cl/ segment A1 has slightly higher calculated crystallization
Nb (28–48) samples from the SMAR (137-3, 143-1, 143-2, pressures than the shallower segment A2 with on average
147-2, 147-3, 149-1, 149-4 and 151-3; Online Resource 4). 385 and 322 MPa, respectively. At Gakkel Ridge, the high-
The SMAR melt inclusions display a generally positive trend est average pressures are found in the Eastern Volcanic Zone
between Cl concentrations (20–168 ppm in Ol, 30–234 ppm (538 MPa), while the Western Volcanic Zone and Sparsely
in Pl) and K (67–2256 ppm), but with more scatter towards Magmatic Zone display lower pressures (340 and 375 MPa,
higher Cl-values than observed in the bulk glasses (Fig. 6a). respectively).
This scatter is not concentration-dependent, and is consider- For the rare samples that contain Cpx (SMAR 147-2,
ably larger than analytical error. Cl/K ratios in the inclusions 149-1, 149-4, 162-1, 162-2), we also calculated pressures
(0.08–0.56) extend mainly to higher values than observed (given in Table  1) using the formulations for Cpx–melt

13
97   Page 10 of 23 Contrib Mineral Petrol (2017) 172:97

Fig. 5  Indicators of hydrothermal Cl addition (Cl/Nb, Cl/K) in basal- Fig. 6  Melt inclusion Cl and K data. a Melt inclusions in Ol (dia-
tic glass plotted against major and trace element data. The analytical monds) and Pl (circles) show a similar trend in Cl and K as the
error is smaller than the symbol sizes. a Cl/K ratios reach higher val- SMAR glasses, but with larger scatter. There is no systematic dif-
ues than previously observed for average slow-spreading ridges, but ference between melt inclusions in Ol or Pl. b Cl/K vs. K for melt
are not as high as observed for Red Sea Rift basalts or fast-spread- inclusions (filled symbols) together with their host glasses (stars; col-
ing ridges. High Cl/K (>  0.1) is not observed for samples with K/ ours distinguish samples). Melt inclusions in minerals hosted in trace
Ti  >  0.22. Numerous b Gakkel Ridge and c SMAR samples have element depleted glasses display both lower and higher Cl/K and K
Cl/Nb above the Cl/Nb mantle ratio. Samples from Gakkel Ridge than their host glasses. Note that the most K-depleted inclusions show
show high Cl/Nb over the complete range of Nb; at the SMAR this is extremely high Cl/K ratios, exceeding the range in their host glasses.
mainly for samples with low Nb c Melt inclusions from trace element-enriched (high K) glass sample
149-1 have lower Cl and K than their host and overlap with concen-
trations of melt inclusions in trace element depleted samples (b)
equilibrium of Putirka et  al. (1996). Melt–phenocryst
equilibrium was assessed by comparing the predicted and
observed values for the Cpx components (cf. Putirka 2008)

13
Contrib Mineral Petrol (2017) 172:97 Page 11 of 23  97

and by comparing the observed and predicted values for KD Salters and Stracke (2004) choose 6.9 and Saal et al. (2002)
(Fe–Mg)cpx–liq (Online Resource 5; Table 1). In terms of give a ratio of 3. However, Palme and O’Neill (2003) cal-
the Cpx components, the calculated and observed composi- culate a relatively high Cl mantle value (30 ppm) compared
tions were within 5% and indicate equilibrium in all sam- to values that were derived from basaltic melts and melt
ples. Samples 147-2 and 162-2 also show equilibrium in inclusions, which are all lower than 7 ppm (Ryabchikov
KD (Fe–Mg)cpx–liq. Samples 149-4 and 162-1 display minor 2001; Saal et al. 2002; Salters and Stracke 2004; Kovalenko
[KD(Fe–Mg)cpx–liq = 0.17–0.20] and sample 149-1 major  et al. 2006; le Roux et al. 2006), while values derived from
[KD(Fe–Mg)cpx–liq = 0.12] disequilibrium in Fe and Mg of mantle minerals are even lower < 0.38 ppm (Urann et al.
the Cpx with the glass, indicating some late-stage modi- 2017). If we assume mantle concentrations of 0.3–0.6 ppm
fication of the glass for these fast-diffusing elements. As for Nb (Hofmann 1988; McDonough and Sun 1995; Palme
the pressure calculations are dependent on variations in the and O’Neill 2003) and a Cl concentration of < 7 ppm, then
Cpx components, which involve slower diffusing elements, mantle Cl/Nb ratios should lie at least below the value of
this minor disequilibrium in Fe–Mg for samples 149-4 and 25.8 proposed by McDonough and Sun (1995). To deter-
162-1 has no significant effect on the resulting pressure cal- mine in more detail the local Cl/Nb in our study areas we
culations. The disequilibrium effect is significant, however, can use the variations we see in our samples. Assuming (a)
for sample 149-1 so we have excluded it from the subse- no fractionation of Cl/Nb during melting (which can be
quent discussion. The four samples in (close) equilibrium excluded due to their similar distribution coefficients) and
yield pressures between 300 and 400 MPa, which are not (b) no degassing of Cl from the magma prior to eruption
consistent with the pressures calculated by COMAGMAT (see below) then interactions of the magma with its environ-
(Table 1). This will be discussed further below. ment should always lead to Cl/Nb increases. In this case, the
magmas with the lowest Cl/Nb represent the least contami-
nated samples and their Cl/Nb should be closest to the local
Discussion (upper limit) mantle Cl/Nb ratio. The evidence for a mix of
sources at both locations may imply that the Cl/Nb and Cl/K
Chlorine contents and Cl/Nb values of the mantle mantle ratios are not constant within an area, which may be
and magma particularly the case for the SMAR ridge samples as HIMU
sources may have higher Cl/K values (Stroncik and Haase
In order to identify Cl addition to the magmas from external 2004; Kendrick et al. 2017). However, Fig. 5b, c, shows a
processes, it is necessary to determine the possible range in very consistent lower Cl/Nb ratio in our samples over a large
local mantle Cl contents (and hence its probable concentra- range of absolute Nb contents at both locations. At Gakkel
tions in mantle-derived magmas). The correlation between Ridge these values span the full range of compositions and
Cl and many incompatible trace elements (particularly those a constant Cl/Nb mantle value is confirmed by the fact that
that have a compatibility close to Cl, for example Nb) and we see, for example, no correlations between Cl/Nb and any
to Sr-, Nd- and Pb-radiogenic isotopes in the case of the trace element ratios indicating variations in continental input
SMAR (Fig.  4b–d) indicates that much of the variation (e.g. Ce/Pb, U/Nb, cf. Hofmann et al. 1986) at the Gakkel
in absolute Cl concentrations between the samples can be Ridge.
explained by source processes. Mantle source variations at In the SMAR samples we see a negative correlation
the SMAR were described by, e.g. Almeev et al. (2008) and between the lowest Cl/Nb values and trace element con-
Hoernle et al. (2011) and were attributed to mixing between centrations as all depleted SMAR basalts (Nb < 4.5) show
a depleted mantle source (with low Cl, Nb and K concentra- higher Cl/Nb ratios (significantly above analytical error, see
tions) and several enriched (HIMU) mantle sources (with “Analytical methods”), while SMAR samples with low Cl/
higher Cl, Nb and K concentrations), while variations in Nb and Nb > 4.5 ppm fall on a line with Cl/Nb of 20 ± 2
mantle source compositions at the Gakkel Ridge were identi- (Fig. 5b). We cannot rule out that this results from mixing
fied by, e.g. Michael et al. (2003) and Goldstein et al. (2008). of a magma from a low Nb—high Cl/Nb (~ 50) source with
We see from Fig. 4f, g that each individual axial locality has magma from a trace element enriched, low Cl/Nb source,
its own inherent Nb/K signature and, by inference, Cl/Nb but this appears unlikely for the following reasons: (a) we
and Cl/K mantle signatures. If we can define these mantle see a range of Cl/Nb values for similar trace element (Nb)
ratios for the individual regions, then Cl/Nb and/or Cl/K concentrations that cannot be explained by mantle varia-
above these values should indicate Cl addition. tions; (b) we see no positive correlation between Cl/Nb and
Cl/Nb (and Cl/K) mantle ratios are not well constrained other indicators of mantle enrichment (e.g. Zr/Hf; Nb/Zr);
and cover a large range in the literature: Palme and O’Neill (c) depleted mantle sources have been shown to have lower
(2003) suggest the highest Cl/Nb of 51, McDonough and Cl/K (and by implication, Cl/Nb) than enriched, particularly
Sun (1995) propose 25.8, le Roux et al. (2006) favour 14, HIMU, mantle sources (Michael and Cornell 1998; Stroncik

13
97   Page 12 of 23 Contrib Mineral Petrol (2017) 172:97

and Haase 2004; Kendrick et al. 2017). Therefore, if the of 0.040 and 0.086 (Fig. 6b, c), which is consistent with
depleted mantle source (low Nb samples) has a different typical Cl/K values of 0.02–0.04 and 0.04–0.08 determined
Cl/Nb ratio, this is expected to be lower than the Cl/Nb of by Stroncik and Haase (2004) for EM (Gakkel Ridge) and
20 ± 2 indicated by the HIMU source samples (high Nb HIMU (SMAR) suites.
values). Instead, the higher Cl/Nb for all SMAR samples Using these Cl/Nb mantle ratios, we can calculate Cl-
with Nb < 4.5 ppm, more likely can be explained by their excess in ppm for each sample based on its Nb or K content
inherently higher sensitivity to Cl contamination due to their (Fig. 7a, b). In contrast to the absolute Cl/Nb ratios, Cl-
low magmatic Cl (the effect of adding a constant amount of excess is independent of the trace element (e.g. Nb) con-
excess Cl to magmas with different Nb concentrations are centrations of the samples, and shows only the addition of
shown as dashed lines on Fig. 5b, c; note that the smooth- Cl to magmas derived from any particular mantle source.
ness of the curves observed are an artefact of plotting the Cl-excess is up to 250 ppm, and ~ 75% of the measured sam-
ratio of an element against the same element). We thus take ples from both the Gakkel Ridge and SMAR show Cl addi-
these values to indicate the local mantle Cl/Nb ratios (ca. tion (Online Resources 2, 3) over a range of compositions,
20 ± 2 at SMAR, 18.5 ± 1.5 at Gakkel). All samples with spreading rates and both comparatively thick and thin crust
Cl/Nb above these local mantle Cl/Nb ratios are taken to (Fig. 7). The amount of Cl-excess is similar for the SMAR
show Cl-excess (defined as Cl content above the expected and Gakkel Ridge, apart from samples from the EVZ that
Cl in the melts based on their Nb content and the local Cl/ show significantly less Cl-excess (Fig. 7a). The SMAR melt
Nb mantle ratio). inclusions show the same extent of Cl-excess as the glasses
Using the correlation regression shown in Fig. 4f, Cl/Nb (Fig. 7b). Both ridges show considerably lower Cl-excess
mantle ratios of 18.5 ± 1.5 and 20 ± 2 for the Gakkel Ridge than basalts from the slow-spreading Red Sea (< 1400 ppm;
and SMAR, respectively, correspond to Cl/K mantle ratios Fig. 7a; cf. van der Zwan et al. 2015).

Fig. 7  Cl-excess (ppm) calcu-


lated by the Cl concentrations
minus the magmatic Cl based
on Cl/Nb mantle values plotted
against major and trace element
data. a Cl-excess is present in
the glasses over various Nb
concentrations and is up to
250 ppm. b The melt inclusions
show the same of Cl-excess as
the glasses. c Cl-excess is not
related to MgO (fractionation),
d ­Na8 (melt degree), e 87Sr/86Sr
(source variation) or f ­CO2
(degassing)

13
Contrib Mineral Petrol (2017) 172:97 Page 13 of 23  97

Chlorine excess and magmatic, seafloor (Fig. 6c) further underlines their xenocrystic origin and the
and hydrothermal processes at SMAR and Gakkel late addition of the enriched component to the system before
Ridge eruption (Fig. 6c). The occurrence of rock fragments identi-
fied by deformation textures in the SMAR samples (Figs. 2,
Magmatic processes such as fractional crystallization and 3) suggests that the magmas assimilated oceanic crust or
partial melting as well as source variations influence the partially solidified crystal mushes. Evidence for assimila-
absolute Cl contents of the melts, but do not affect the Cl/ tion of country rocks at Gakkel comes from U-series data
Nb ratios and hence cannot cause the observed Cl-excess. (Elkins et al. 2014).
This is additionally confirmed by the lack of a correlation To determine whether the Cl-excess in our glasses comes
between Cl-excess and indicators of fractionation (MgO) from assimilation of country rock during ascent we can use
or degree of melting (­ Na8) and source variations (87Sr/86Sr; the ­H2O/Cl ratios of the samples, as this ratio is different in
Fig. 7c–e). Degassing of the samples is indicated by the seawater, altered rock and hydrothermal brines. When we
­CO2/Nb of the samples of mostly < 200 and the C ­ O2/Ba plot ­H2O/Cl against K/Cl (following Kendrick et al. 2013;
of < 60 that are significantly lower than values of 239 or Fig. 8), we see that our samples follow a mixing line from
283 for C­ O2/Nb and 105 for C­ O2/Ba for undegassed magma mantle values to a contaminant with low ­H2O/Cl (< 9). Both
derived by Saal et  al. (2002) and Michael and Graham hydrothermally altered lithosphere and seawater have high
(2015). Degassing of ­CO2 can also be seen in a plot of C ­ O2 ­H2O/Cl ratios (> 16 and ~ 50; see Ito et al. 1983; Bach
against ­H2O compared to isobars of equilibration eruption et al. 2003; Sano et al. 2008; Barnes and Cisneros 2012)
pressures where we see degassing trends for ­CO2 towards the and would introduce H ­ 2O to the magma. The good cor-
eruption pressures of the samples (Online Resource 6). ­H2O relation between ­H2O and Ce (maximum 0.22 wt% varia-
shows, however, no degassing trends and is undersaturated tion for a given Ce content per sample suite; Fig. 4e), the
compared to the isobars, implying no ­H2O degassing took undersaturation of the magmas in H ­ 2O (Online Resource
place, consistent with the good ­H2O/Ce correlation. Also 6), and H ­ 2O/Ce values of < 350, all indicate no significant
variations in Cl due to Cl removal by degassing can be ruled ­H2O addition (see also Michael 1995; le Roux et al. 2006;
out as there is no correlation between C ­ O2/Nb or ­CO2/Ba Wanless et al. 2010, 2011; van der Zwan et al. 2015). The
and Cl-excess (Fig. 7f). Cl-excess we observe cannot, therefore, be explained by the
As melt inclusions display the same Cl-excess as the bulk addition of either seawater or bulk altered lithosphere.
glasses (Figs. 5a, 6, 7a, b) and Cl-excess is homogeneous A Cl-rich phase (that adds up to 250 ppm Cl) with low H ­ 2O/
across multiple glass chips of the same sample and also not Cl can be introduced by assimilation of 0.25–0.08% fluids
correlated to indicators of surficial alteration e.g. U, Ba (e.g. with NaCl contents of 16 to > 50%, respectively (brines, e.g.
Alt et al. 1986; Alt and Teagle 2003; Schramm et al. 2005; Michael and Schilling 1989; Kendrick et al. 2013; van der
Augustin et al. 2008), Cl addition by surficial processes (sea- Zwan et al. 2015; Fig. 8). Such fluids are generated, particu-
floor alteration or syn-eruptive magma–seawater interaction, larly at high temperatures and pressures, by phase separa-
e.g. Hart et al. 1974; Soule et al. 2006) can be ruled out tion of hydrothermal fluids (Bischoff and Rosenbauer 1987;
as playing a significant role in generating the Cl-excess we
observe. Instead, the source of the Cl-enrichment must lie
deeper in the magmatic system, before eruption but during
crystal growth and magma homogenization.
The assimilation of hydrothermally altered oceanic crust
was shown to lead to Cl-excess in glasses at other locations
(cf. Bédard 1991; Bédard and Hébert 1996). Furthermore,
the global lack of Cpx in most MORB magma despite a
positive correlation between Ca and Mg has been explained
by assimilation of (lower) crustal rocks (e.g. Kvassnes and
Grove 2008; Lissenberg and Dick 2008; Godard et al. 2009;
Drouin et al. 2009) and melt–rock reactions (syntexis; Béd-
ard et al. 2000 and references therein; Kelemen et al. 1995).
The petrology of our SMAR samples shows that they have a Fig. 8  The relation of the K/Cl and the H­ 2O/Cl contents of the sam-
complex history of melt/rock interactions, with phenocryst ples compared to their mantle values and potential contaminants
rim/glass disequilibrium (Fig. 3b), complex or reverse zon- (seawater, altered oceanic crust, brines). Salinity of the brines with
10–50 wt% salts and mantle data are after Kendrick et al. (2017), but
ing and dissolution features (Fig. 2) all suggesting magma
with the local K/Cl mantle values as derived in this paper. Altered
mixing and xenocryst inheritance. The presence of depleted oceanic crust data are from Ito et al. (1983), Bach et al. (2003), Sano
melt inclusions in phenocrysts from enriched magmas et al. (2008), Barnes and Cisneros (2012). See text for discussion

13
97   Page 14 of 23 Contrib Mineral Petrol (2017) 172:97

Fournier 1987; Berndt and Seyfried 1990). Hydrothermal cotectic as the defining final event in the magmatic evolution
brines could potentially be added directly to the magmas is probably incorrect. Petrographically and geochemically
if fluids intrude through the conductive cracking front of our samples show signs of mineral–melt chemical disequi-
the magmas (e.g. Kendrick et al. 2013); the petrographic librium (Fig. 3), assimilation of lithic xenoliths (Fig. 2)
evidence for assimilation of country rocks in both sample and magma mixing (Figs. 3, 6). Moreover, the lack of Cpx
sets argues, however, against such a simple incorporation phenocrysts in most of the samples is inconsistent with the
of fluids. Assimilation of hydrothermally altered rocks with assumption that the melts are at a multiply-saturated cotectic
brines present in the pore-spaces (Coogan et al. 2003) or as with Ol, Pl and Cpx.
fluid inclusions in hydrothermally formed minerals, would Melt–rock reactions in the lower lithosphere may signifi-
lead to Cl-excess but little H
­ 2O addition. This mechanism for cantly change the compositions of both the cumulates (e.g.
Cl-contamination has also been suggested for samples at the Meyer et al. 1989; Bédard 1991; Bédard and Hébert 1996;
boundary between the magmatic and hydrothermal systems Bédard et al. 2000; Coogan et al. 2000; Dick et al. 2002;
from IODP Site U1256 (fast-spreading EPR), where Cl was Ridley et al. 2006; Grimes et al. 2008) and the melt. As
found to be present in fluid inclusions or pore fluids (Zhang we have no precise information on the original melt and
et al. 2017) and stoping of altered crust was thought to be the contaminant compositions, the extent of assimilation can-
dominant Cl enrichment process (Fischer et al. 2016). Alter- not be determined straightforwardly by backward-modelling
natively, Cl could be added by partial melting (anatexis) of although Kvassnes and Grove (2008) and Lissenberg and
the hydrothermally altered lithosphere and particularly by Dick (2008) have shown, using experiments and forward
the breakdown of Cl-rich hydrothermally formed minerals modelling, that assimilation of various types of oceanic
such as amphiboles (Barnes and Cisneros 2012); these min- crust has a significant effect on the composition of the melts
erals melt preferentially due to a lower solidus than the fresh and their ­Al2O3–CaO–Si–MgO relations. They concluded
rock, and partition Cl much more strongly than ­H2O into the that any method that uses the composition of a melt (glass)
melt, producing a low ­H2O/Cl contaminant (Michael and alone to determine, e.g. crystallization pressure or temper-
Schilling 1989; Kent et al. 1999; France et al. 2010, 2014; ature from a melt can be erroneous. Lissenberg and Dick
Wanless et al. 2010, 2011). This process has been proposed (2008) further showed that calculated pressures increase
for some basaltic to dacite lavas on the EPR based on their with increasing assimilation of lower crustal troctolite (up
major and trace element data (France et al. 2010, 2014; Wan- to 25%) and that the full range of crystallization pressures
less et al. 2010, 2011) although it was excluded for the rocks observed in MORB (100–800 MPa) can be achieved by reac-
from IODP Site U1256 which apparently reacted under dry tion of melts with the lower crust. This is consistent with the
conditions without the influence of amphibole (Fischer et al. basalt data from the SMAR, where many of the samples that
2016; Erdmann et al. 2017). Since there is no correlation in display high COMAGMAT pressures (> 300 MPa; Table 1)
our samples between Cl-excess and any of the other meas- show Cl-excess implying they have assimilated altered crust.
ured elements (Fig. 7; Online Resources 2, 3) that could be As all samples with Cl-excess likely have modified major
indicative for any of these processes, we cannot distinguish, element compositions due to assimilation, pressure calcula-
with our data, between the assimilation of altered country tions based on major elements cannot be used to determine
rock containing interstitial brines or brine in fluid inclusions the depth of Cl addition.
and partial melt of (potentially amphibole-bearing) altered Pressure calculations based on Cpx–melt element
lithosphere. We therefore group all these potential deep Cl exchange (Putirka 2008) can give a better control on the
contamination processes under the term ‘assimilation of crystallization pressures, as the degree of equilibrium
hydrothermally altered lithosphere’. between glass and Cpx can be assessed and, by using the
Cpx rim compositions, only equilibrium in the last phase of
Crystallization depths and the validity of the pressure crystallization is required, thus the pressures are not affected
calculations by earlier magma modifications. A statistically robust
assessment of the relationship between Cl-excess and crys-
The discrepancies between the pressure of last equilibrium tallization pressures is hindered by the lack of Cpx in most
calculated using COMAGMAT (0–840  MPa) and those (MORB) samples, but our samples with Cpx in equilibrium
derived from the Putirka et al. (1996) Cpx–melt equilib- give geologically relevant depths.
rium model (300–400 MPa) indicate that at least one type
of pressure estimate is inaccurate. The COMAGMAT pres- The depth of assimilation of hydrothermally altered
sures are representative for melt-based pressure calculations lithosphere
(see Online Resource 7 for discussion), but the assumption,
inherent in all phase-based pressure calculations, of a sim- The SMAR Cpx–melt pressures (after Putirka et al. 1996) all
ple crystallization history and equilibrium at the Ol–Pl–Cpx lie in a relatively restricted range (316–403 ± 150 MPa ≈ 

13
Contrib Mineral Petrol (2017) 172:97 Page 15 of 23  97

10–13 ± 5 km; Table 1). For a crustal thickness of ~ 11 km many Ol grains are xenocrystic, they must have crystal-
on segment A2 (Bruguier et al. 2003) where these samples lized from earlier batches of melt that already assimilated
originate, these pressures imply that the last episode of crys- Cl as shown by the Cl-excess of the melt inclusions in
tallization occurred at lower crustal depths or close to the the Ol. Hence, the various amounts of Cl-excess observed
Moho. The Cl-enriched melt inclusions in Ol and Pl imply in the glass and melt inclusions from one rock probably
an increase of Cl in the magma prior to crystallization and represent multiple generations of hydrothermal alteration,
inclusion formation. The complete equilibrium observed for assimilation of crust and crystallization, rather than pro-
samples 147-2 and 162-2 between the Cpx phenocrysts and cesses occurring all to the same magma batch. Production
melt show no substantial modification of the melt subse- of the crust by multiple cycles of crystallization was also
quent to Cpx crystallization. Thus Cpx crystallized together suggested by Grimes et al. (2008) and Lissenberg et al.
or after the Ol and Pl that host the Cl-contaminated melt (2009), who found a range of ages in gabbros from the
inclusions. This places a strong limit on the minimum depth Mid-Atlantic Ridge.
of hydrothermal circulation in this region of ~ 10 km. We We see very similar degrees of Cl-excess in all regions
note that not all samples that display lower crustal Cpx crys- studied and the homogeneous distribution of Cl in the bulk-
tallization pressures display Cl-excess, indicating that not all rock glasses from both the SMAR and Gakkel Ridge shows
assimilated crust is altered or that assimilation of crust is not extensive homogenization after assimilation, which requires
the only method of magma cooling at depth. magma reservoir processes. Therefore, also at Gakkel Ridge,
Also the investigated samples that show a slight modifica- Cl cannot simply have been added during ascent, imply-
tion of the host melt after crystallization of the Cpx of the ing that hydrothermal circulation must reach the levels of
faster diffusing elements by minor disequilibrium in Fe–Mg magma pockets at depth. At the ultraslow-spreading Gakkel
between the Cpx and glass, but with Cpx components that Ridge, the thin lithosphere (Reid and Jackson 1981; Dick
are in equilibrium, host melt inclusions with Cl-excess indi- et al. 2003; Montési and Behn 2007) will be inherently more
cating deep assimilation of brines (Table 1). Disequilibria strongly cooled by conduction, which promotes deep crys-
between melts and several Ol rims imply assimilation or tallization and so magma reservoirs are expected to form
mixing processes shortly before eruption, as most Ol in deeper there (Rubin and Sinton 2007; Morgan and Chen
disequilibrium do not show any evidence of diffusive re- 1993). Both deep pooling of melt (Shaw et al. 2010) as well
equilibration that would have led to the formation of obvious as melt–rock interactions at depths < 15 km (Elkins et al.
diffusion profiles within a few days at magmatic tempera- 2014) have been suggested at Gakkel Ridge. This evidence
tures (cf. Klügel 1998; Shaw and Klügel 2002). Even the one and the consistent level of Cl-excess between the ridges
observed diffusion profile with a diffusion length of < 50 µm implies that deep hydrothermal circulation is also occur-
(sample 143-2; Fig. 3a), indicates a short residence time ring there. Crystallization and crustal assimilation in magma
of the Ol in the melt of maximum a few weeks (cf. Shaw bodies in the lower lithosphere at (ultra)slow spreading
and Klügel 2002). The lack of diffusion profiles in the Ol ridges is consistent with the evidence of gabbroic samples
(Rutherford 2008) shows that the rising of magma was rela- and melt–rock reactions in the lower crust and mantle found
tively fast (days to a few weeks) and therefore Cl addition by at the South-West Indian Ridge and the MAR 30°N (e.g.
assimilation of crust is less likely to happen during ascent Dick et al. 2000; Expedition Scientific Party 2005).
in the upper crust. This is in agreement with what would be Deep circulation of water can be facilitated by deep faults-
expected at slow-spreading ridges, where the cooler upper pathways. Data for teleseismic earthquakes (1960 to present;
crust (Reid and Jackson 1981; Dick et al. 2003; Montési International Seismological Centre 2011) record events at
and Behn 2007) requires that magma rises fast in order to depths of > 10 km for both study areas. Although the depth
not crystallize. estimate of these earthquakes is not very precise (Engdahl
Clinopyroxene phenocrysts were only found at SMAR et al. 1998), the > 3 magnitude of these earthquakes (Fig. 1)
A2, where the anomalously thick crust probably favours indicates that the faults causing them must be large and thus
the early crystallization of clinopyroxene. At slow-spread- have the potential to extent to larger depths. Also, micro-
ing ridges with typical or lower crustal thickness without seismicity on SMAR segment A1 down to 2–4 km within
clinopyroxene phenocryst (SMAR segment A1 and Gak- the 5-km-thick crust, indicates that faults extend to lower
kel Ridge) crystallization and assimilation depths cannot crustal depths (Grevemeyer et al. 2013). Deep faults, down
be directly determined. Nevertheless, the ranges in Cl, K, to the lower crust, were found in gabbros at slow-spreading
Cl/K and Cl-excess observed in melt inclusions in indi- ridges, e.g. by Harper (1985) Cannat et al. (1991) and Mével
vidual samples also from SMAR segment A1 imply that and Cannat (1991), while Bach et al. (2004) showed that
they were formed from both different mantle melts and also serpentinites that may be present in the lithosphere of
melts which had experienced varying degrees of assimi- ultraslow-spreading ridges can break and alter at depth. In
lation before extensive crystallization (Figs. 6, 7b). As addition, Hasenclaver et al. (2014) showed that the physical

13
97   Page 16 of 23 Contrib Mineral Petrol (2017) 172:97

process of deep hydrothermal circulation can be modelled higher average potential to host hydrothermal systems due
at greater depths. to a higher magmatic activity, shown by detailed studies
In conclusion, hydrothermal circulation is not restricted on the relation between hydrothermal occurrences and rift
to the upper crust and the assimilation depth is therefore geomorphology (Fouquet 1997; Anderson et al. 2017). The
not a limiting factor on Cl-addition to rising magmas as axial high of segment A2 was earlier interpreted to be a
suggested by Michael and Cornell (1998). The difference remnant of past high magmatic activity (Devey et al. 2010)
between our conclusion and those of previous work is prob- and there are currently no signs of hydrothermal activity
ably related to analytical precision, which needs to be high (Devey et al. 2010; Haase et al. 2016). Therefore, we take
to distinguish between small, but significant variations in the high Cl-excess in basalts at the axial highs of A2 and in
Cl/Nb and to identify appropriate Cl/Nb mantle values and the north of A1 to indicate that alteration of the crust took
Cl-excess in slow-spreading-ridge magmas. place there previously by ancient, now inactive hydrothermal
systems, which makes these areas good targets for the search
Chlorine excess in lavas and sites of hydrothermal for extinct hydrothermal deposits.
activity At Gakkel Ridge, samples collected within 10 km of the
Aurora hydrothermal field at 6°20′W and plume sites at
To study the spatial relationship between Cl-excess in lavas 1°45′W and 37°E display high Cl-excess (> 25 ppm; Fig. 9b)
and hydrothermal vent sites, we compared the sampling consistent with the observations at the SMAR. Plume sites
locations of basalts showing Cl-excess with those of known at 2°10′E, 7°30′E and 43°10′E display samples with high
active hydrothermal vent sites and indirect indications for Cl-excess within the larger vicinity of 25  km, which is
hydrothermal activity (e.g. plume signatures; Fig. 9). At both likely related to the highly uncertain locations for the origin
the SMAR and Gakkel Ridge there are distinct areas with of most of the hydrothermal plumes (Baker et al. 2004).
multiple lavas containing Cl-excess. These areas are most The only exception among our samples are basalts that do
apparent on the SMAR (Fig. 9a), where, within 10 km of a not exhibit Cl-excess collected < 10 km from the present-
known hydrothermal site or plume, samples with Cl-excess day hydrothermal active site at 85°E (Fig. 8b; Baker et al.
are always seen, which are also some of the highest val- 2004; Pontbriand et al. 2012). But in detail, these sampled
ues observed along the ridge. High Cl-excess is also seen glasses come from visibly older lava flows than the young
at the axial high of segment A2 (8°57′S) and the relatively glassy basalts, believed to have erupted in 1999 that the pre-
shallow area in the north of A1 (7°51′S; Fig. 9a). Axial sent hydrothermal venting was linked to (Pontbriand et al.
topographic highs and domes are volcanic structures with a 2012). We therefore suggest that the 85°E system is a young

Fig. 9  Locations of high
Cl-excess samples (diamonds)
and samples without Cl-excess
(circles) and hydrothermal sites
at the a SMAR and b Gakkel
Ridge. Note in the vicinity of a
known hydrothermal site there
are always samples taken with
elevated Cl-excess (> 25 ppm)
with the exception of 85°E at
Gakkel Ridge, where independ-
ent evidence suggests that the
hydrothermal activity is prob-
ably younger than the magmas
sampled (for discussion see
text). The axial high at SMAR
segment A2 (8°57′S) yielded
Cl-excess samples but lacks any
recent hydrothermal activity,
which suggests that the high Cl-
excess results from former, now
inactive hydrothermal activity

13
Contrib Mineral Petrol (2017) 172:97 Page 17 of 23  97

hydrothermal system, that was established after the eruption fast-spreading ridges or ridges with a thick crust, contrary
of the here measured samples. to the conclusions of Michael and Cornell (1998), but also
In summary, Cl-excess in erupted lavas always indicate occurs in several ultraslow- and slow-spreading ridges
that they have interacted with altered crust implying active that show a range of morphologies and thicknesses. Nev-
or former hydrothermal circulation in the vicinity; the large ertheless, differences in the amounts of Cl-contamination
amount of samples with Cl-excess show that hydrothermal between fast- and slow-spreading ridges are clearly evident.
alteration is widespread. Basalts with the highest Cl-excess Following Stroncik and Niedermann (2016) we assembled
(> 25 ppm) are related to places that display active or fossil Cl and Cl/K data from the PetDB database (Lehnert et al.
hydrothermal venting in the vicinity and as such represents 2000) and used them to assess degrees of Cl-excess globally.
a guide that can aid to identify MOR segments with hydro- Absolute MORB Cl-concentrations at slow-spreading ridges
thermal deposits. are overall considerably lower (only 14% has > 200 ppm
Cl) than at fast- or medium-spreading ridges (36 and 31%
Differences in Cl contamination between slow‑ has > 200 ppm Cl; Fig. 10a). Using an upper value of 0.09
and fast‑spreading ridges for mantle Cl/K implies that at least 25% of the samples at
slow-spreading ridges show Cl-excess (Fig. 10b), consider-
Our data show that Cl-contamination by assimilation of ably less than observed at fast- and medium-spreading ridges
hydrothermally altered lithosphere is not restricted to (69 and 79%). The maximum Cl-excess on slow-spreading

Fig. 10  Chlorine concentrations at different ridge types. a Frequency value of 0.09 (dashed line). c Comparing samples with definite Cl-
plots of Cl-contents and b Cl/K for slow-, medium- and fast-spread- excess with hydrothermal sites on the most studied, slow-spreading
ing ridges. Slow-spreading ridges show on average lower Cl-contents Norther Mid-Atlantic Ridge (20°–65°N) shows that Cl-excess is pre-
and Cl/K than medium- and fast-spreading ridges, nevertheless 23.9% sent along the whole ridge and that where data are present hydrother-
of the slow-spreading MORB show Cl/K above a upper mantle limit mal sites are mostly associated with Cl-excess samples close by

13
97   Page 18 of 23 Contrib Mineral Petrol (2017) 172:97

ridges in the PetDB database overlaps with our study and alteration processes are more localized, explaining
(< 550 ppm, but mostly < 250 ppm) and is clearly lower than that not all assimilated rock is altered at slow-spreading
at fast- and medium-spreading ridges (mostly < 1000 ppm). ridges leading to more variable amounts of Cl-excess in the
The fact that only ca. 25% of slow-spreading magmas in samples. The large permanent magma lenses beneath fast-
PetDB are classified as showing Cl-excess compared to spreading ridges provide the opportunity for more extensive
our 75% estimate for Gakkel and SMAR shows that using mixing of melts and assimilation of the crust over a larger
a global mantle Cl/K of 0.09 underestimates the samples scale, leading to a common chemical signature of assimila-
with Cl-excess at slow-spreading ridges and emphasizes the tion of hydrothermally altered crust (cf. Saal et al. 2002;
need for both high-precision Cl determinations and good Kendrick et al. 2013).
estimates of local mantle Cl/K values in these inherently Cl-
poor magmas. Cl-excess in the PetDB data is not restricted Global implications of hydrothermal Cl contamination
to specific sections of slow-spreading ridges, e.g. with thick
crust as suggested by Michael and Cornell (1998), as for The depth of hydrothermal circulation is important for the
example can be seen from the well-studied slow-spreading cooling of the lithosphere. With our data, we can demon-
Northern Mid-Atlantic Ridge (Fig. 10c). strate that there is a cooling mechanism by circulation of
Hence, Cl addition by hydrothermal contamination is hydrothermal fluids down to the lower crust that can facili-
present on all types of ridges, but shows a difference in the tate deep crystallization at slow-spreading ridges. This cir-
amount of Cl-excess in basalts from the two end-member culation most likely happens at all low spreading rates. This
type of ridges. While Cl-excess at fast-spreading ridges is means crust is not merely formed at shallow levels but that
high and present in most samples (e.g. Michael and Cornell lower crustal and lithospheric mantle magma chamber pro-
1998; Saal et al. 2002; le Roux et al. 2006; Kendrick et al. cesses may be common at slow-spreading ridges and that
2013), Cl-excess at slow-spreading ridges is more variable oceanic crust can form in situ at all depths (cf. Kelemen
and significantly lower (Fig. 5a; Michael and Cornell 1998; et al. 1997; Kelemen and Aharonov 1998; Lissenberg et al.
van der Zwan et al. 2015; this paper). The lower absolute 2004; Grimes et al. 2008).
intensity of Cl-contamination in slow-spreading magmas Hydrothermal circulation throughout the newly formed
implies either a lower degree of assimilation and/or a lower lithosphere can increase the Cl budget and to a smaller
degree of lithospheric alteration. The less differentiated extent the ­H2O budget of an altered oceanic slab and thus
nature of magmas from slow- compared to fast-spreading the potential of Cl (and ­H2O) to subduct, although we cannot
ridges (higher average Mg#; Rubin and Sinton 2007) sug- determine with the present data the Cl content and distribu-
gest that the magmas at slow-spreading ridges on average tion in the altered plate and so cannot give a realistic esti-
underwent less fractional crystallization and therefore mate of its total Cl content. This has important consequences
have lost less heat during passage through the lithosphere. for global water and halogen budgets (see e.g. Kendrick et al.
Whether that results in less extensive hydrothermal circula- 2017). Although most other elements occur in too low con-
tion and alteration or less assimilation of country rock, the centrations in hydrothermal fluids to significantly change the
net effect will be a lower average Cl addition into the slow- composition of the deeper lithosphere, H- and S-isotopes, B
spread magmas. A decrease of hydrothermal alteration with and noble gasses may be significantly changed (Kent et al.
depth (e.g. Dick et al. 2000; Coogan 2003) may also play a 1999; Gillis et al. 2003; Kendrick et al. 2013; Labidi et al.
role in the lower Cl-excesses at slow-spreading ridges. The 2014; Stroncik and Niedermann 2016). Any study of those
lower crustal crystallization depths shown here and deep elements and isotopes should take into account the effect of
assimilation of less altered rock will result in comparatively hydrothermal circulation and assimilation of hydrothermally
minor Cl addition compared to shallower magma chambers altered rocks. In addition, melts that show Cl-excess at slow-
with chronic hydrothermal systems at fast-spreading ridges spreading ridges likely interacted with the lithosphere and
(< 3 km; le Roux et al. 2006; Kendrick et al. 2013) that care should be taken when using their major element data
assimilated more strongly altered crust. If less cooled by for e.g. pressure or melt degree calculations.
hydrothermal circulation, this deeper lithosphere needs to
conductively loose its heat, consistent with thermodynamic
models and inferred temperature gradients that indicate more Conclusion
effective conductive cooling at slow-spreading ridges (Reid
and Jackson 1981; Niu and Hekinian 1997; Montési and Novel high-precision Cl measurements of basaltic glasses
Behn 2007). Slow-spreading ridges display a much higher and melt inclusions in Ol and Pl from the slow- and ultra-
compositional variety with heterogeneous amounts of altera- slow-spreading Southern Mid-Atlantic Ridge and Gakkel
tion (e.g. Dick et al. 2000) and magma bodies are probably Ridge show up to 75% of these samples to have been affected
small and ephemeral, which implies that both assimilation by Cl-contamination, raising their Cl/Nb and Cl/K ratios

13
Contrib Mineral Petrol (2017) 172:97 Page 19 of 23  97

above mantle values, which cannot be explained by mag- References


matic processes such as degassing, melting, fractionation
or source variations, or by seafloor processes. This contami- Almeev RR, Holtz F, Koepke J, Ariskin AA (2007a) The effect of
nation most likely occurs by the addition of hydrothermal minor H2O content on crystallisation in MORB: experiments,
model, applications. In: Goldschmidt 2007 abstracts Geochim
brines to the magmas during the assimilation of hydrother- Cosmochim Acta vol 71 (15S1). p A15
mally altered country rock. Cl-contamination is shown to Almeev RR, Holtz F, Koepke J, Parat F, Botcharnikov RE (2007b) The
be a common process at (ultra-) slow-spreading ridges, in effect of H
­ 2O on olivine crystallization in MORB: experimental
contrast to conclusions drawn from early studies based on calibration at 200 MPa. Am Miner 92(4):670–674
Almeev R, Holtz F, Koepke J, Haase K, Devey C (2008) Depths of
lower precision Cl data. The compositional effects of crus- partial crystallization of ­H2O-bearing MORB: phase equilibria
tal assimilation lead to disequilibrium and phase-equilib- simulations of basalts at the MAR near Ascension Island (7–11
ria-based estimates of magma fractionation depths, such as S). J Petrol 49(1):25–45
those provided by COMAGMAT, to be erroneous. Pressures Almeev RR, Holtz F, Koepke J, Parat F (2012) Experimental calibra-
tion of the effect of ­H2O on plagioclase crystallization in basaltic
calculated from Cpx–melt pairs are more reliable as equi- melt at 200 MPa. Am Miner 97(7):1234–1240
librium can be assessed and yield Cl-contamination depths Alt JC, Bach W (2006) Oxygen isotope composition of a section of
(and thus minimum depths to which hydrothermal circula- lower oceanic crust, ODP hole 735B. Geochem Geophys Geosyst
tion penetrates) of at least 10 km for SMAR segment A2. 7(12):Q12008. https://doi.org/10.1029/2006GC001385
Alt JC, Teagle DA (2003) Hydrothermal alteration of upper oceanic
This is near the base of the lower crust in this studied region. crust formed at a fast-spreading ridge: mineral, chemical, and iso-
Melt inclusions with Cl-excess show assimilation of hydro- topic evidence from ODP Site 801. Chem Geol 201(3):191–211
thermal crust in magma chambers before crystallization at Alt JC, Honnorez J, Laverne C, Emmermann R (1986) Hydrothermal
depth in magma reservoirs. Deep cooling by hydrothermal alteration of a 1 km section through the upper oceanic crust,
Deep Sea Drilling Project Hole 504B: mineralogy, chemistry and
circulation and crustal assimilation is consistent with inde- evolution of seawater–basalt interactions. J Geophys Res Solid
pendent earlier evidence for lower crustal crystallization and Earth 91(B10):10309–10335
melt–rock reactions and for extensive and deep faulting at Anderson MO, Chadwick WW, Hannington MD, Merle SG, Resing
slow-spreading ridges. JA, Baker ET, Butterfield DA, Walker SL, Augustin N (2017)
Geological interpretation of volcanism and segmentation of the
All MORB basalts that display Cl-excess were at some Mariana back-are spreading center between 12.7°N and 18.3°N.
point in time affected by hydrothermal fluids, while the high- Geochem Geophys Geosyst 18(6):2240–2274
est observed values here are associated with hydrothermal Ariskin AA, Barmina GS (2004) COMAGMAT: development of a
magma crystallization model and its petrological applications.
vent fields. Although globally both fast- and slow-spread-
Geochem Int 42(1):S1–S157
ing ridges display a similar proportion of samples with Cl- Augustin N, Lackschewitz K, Kuhn T, Devey CW (2008) Mineral-
excess, slow-spreading ridges show both lower absolute Cl ogical and chemical mass changes in mafic and ultramafic rocks
concentrations and maximum degrees of Cl-excess. This from the Logatchev hydrothermal field (MAR 15 N). Mar Geol
256(1):18–29
may result from either less alteration of the crust or less
Bach W, Peucker-Ehrenbrink B, Hart SR, Blusztajn JS (2003) Geo-
interaction of the magmas with the crust at slow-spreading chemistry of hydrothermally altered oceanic crust: DSDP/ODP
ridges. The generally less fractionated nature of slow-spread- Hole 504B–Implications for seawater-crust exchange budgets and
ing magmas provides support for the latter explanation— Sr- and Pb-isotopic evolution of the mantle. Geochem Geophys
Geosyst 4(3). https://doi.org/10.1029/2002GC000419
the former explanation would require a larger proportion
Bach W, Garrido CJ, Paulick H, Harvey J, Rosner M (2004) Sea-
of conductive cooling of the lower crust as spreading rates water–peridotite interactions: First insights from ODP Leg
decrease. The detection of the effects of deep-seated hydro- 209, MAR 15 N. Geochem Geophys Geosyst 5(9). https://doi.
thermal circulation in slow-spreading crust has implications org/10.1029/2004GC000744
Baker ET, Edmonds HN, Michael PJ, Bach W, Dick HJB, Snow JE,
for the volatile and halogen budget of the oceanic plates and
Walker SL, Banerjee NR, Langmuir CH (2004) Hydrother-
this needs to be taken into account when assessing the Cl, mal venting in magma deserts: the ultraslow-spreading Gak-
­H2O, H-, S-, B- and noble gas isotopic signatures of global kel and Southwest Indian ridges. Geochem Geophys Geosyst
geochemical cycles. 5(8):Q08002. https://doi.org/10.1029/2004GC000712
Barnes JD, Cisneros M (2012) Mineralogical control on the chlorine
isotope composition of altered oceanic crust. Chem Geol 326–
Acknowledgements  We are very grateful to Mario Thöner for the
327:51–60. https://doi.org/10.1016/j.chemgeo.2012.07.022
extensive technical assistance at the EMP and to Dagmar Rau for the
Bédard JH (1991) Cumulate recycling and crustal evolution in the Bay
technical assistance at the LA-ICP-MS. Further, we like to thank Jan
of Islands ophiolite. J Geol 2:225–249
Fietzke (all GEOMAR) for the help with the modification of the Cl
Bédard JH, Hébert R (1996) The lower crust of the Bay of Islands
measurement method for the melt inclusions. The suggestion of three
ophiolite, Canada: petrology, mineralogy, and the importance of
anonymous reviewers and editorial handling by Jochen Hoefs was
syntexis in magmatic differentiation in ophiolites and at ocean
greatly appreciated. We acknowledge generous financial support from
ridges. J Geophys Res Solid Earth 101(B11):25105–25124.
the Jeddah Transect Project between King Abdulaziz University and
https://doi.org/10.1029/96JB01343
Helmholtz-Center for Ocean Research GEOMAR that was funded by
Bédard JH, Hebert R, Berclaz A, Varfalvy V (2000) Syntexis and the
King Abdulaziz University (KAU) Jeddah, Saudi Arabia, under Grant
genesis of lower oceanic crust. In: Dilek Y (ed) Ophiolites and
no. (T-065/430).

13
97   Page 20 of 23 Contrib Mineral Petrol (2017) 172:97

oceanic crust: new insights from field studies and the Ocean Meyer PS, Miller DJ, Naslund HR, Niu Y-L, Robinson PT, Snow
Drilling Program. Geological Society of America, pp 105–120 J, Stephen RA, Trimby PW, Worm H-U, Yoshinobu A (2000) A
Behrens H, Misiti V, Freda C, Vetere F, Botcharnikov RE, Scarlato long in situ section of the lower ocean crust: results of ODP Leg
P (2009) Solubility of ­H2O and ­CO2 in ultrapotassic melts at 176 drilling at the Southwest Indian ridge. Earth Planet Sci Lett
1200 and 1250 °C and pressure from 50 to 500 MPa. Am Miner 179(1):31–51. https://doi.org/10.1016/S0012-821X(00)00102-3
94(1):105–120 Dick H, Ozawa K, Meyer P, Niu Y, Robinson P, Constantin M, Hebert
Berndt ME, Seyfried WE Jr (1990) Boron, bromine, and other trace R, Maeda J, Natland J, Hirth G (2002) Primary silicate min-
elements as clues to the fate of chlorine in mid-ocean ridge vent eral chemistry of a 1.5-km section of very slow spreading lower
fluids. Geochim Cosmochim Acta 54(8):2235–2245. https://doi. ocean crust: ODP Hole 735B, Southwest Indian ridge. Proc
org/10.1016/0016-7037(90)90048-P Ocean Drill Program Sci Results 176:1–61
Berry AJ, O’Neill HStC, Rowe MC, Moselmans JFW, Rivard C (2017) Dick HJB, Lin J, Schouten H (2003) An ultraslow-spreading class of
The oxidation state of iron in basaltic glasses. In: Goldschmidt ocean ridge. Nature 426(6965):405–412
2017 abstracts. Nr 327 Drouin M, Godard M, Ildefonse B, Bruguier O, Garrido CJ (2009)
Bézos A, Humler E (2005) The F ­ e3+/ΣFe ratios of MORB glasses and Geochemical and petrographic evidence for magmatic impregna-
their implications for mantle melting. Geochim Cosmochim Acta tion in the oceanic lithosphere at Atlantis Massif, Mid-Atlantic
69(3):711–725 Ridge (IODP Hole U1309D, 30°N). Chem Geol 264(1–4):71–88.
Bischoff JL, Rosenbauer RJ (1987) Phase separation in seafloor geo- https://doi.org/10.1016/j.chemgeo.2009.02.013
thermal systems; an experimental study of the effects on metal Dunn RA, Toomey DR, Solomon SC (2000) Three-dimensional seis-
transport. Am J Sci 287(10):953–978 mic structure and physical properties of the crust and shallow
Bruguier N, Minshull T, Brozena J (2003) Morphology and tectonics mantle beneath the East Pacific Rise at 9°30′N. J Geophys Res
of the Mid-Atlantic Ridge, 7°–12°S. J Geophys Res Solid Earth Solid Earth 105(B10):23537–23555
108(B2) Edmonds HN, Michael PJ, Baker ET, Connelly DP, Snow JE, Lang-
Cannat M (1996) How thick is the magmatic crust at slow spreading muir CH, Dick HJB, Muhe R, German CR, Graham DW
oceanic ridges? J Geophys Res Solid Earth 101(B2):2847–2857 (2003) Discovery of abundant hydrothermal venting on the
Cannat M, Mével C, Stakes D (1991) Stretching of the deep crust at ultraslow-spreading Gakkel Ridge in the Arctic Ocean. Nature
the slow-spreading Southwest Indian ridge. Tectonophysics 421(6920):252–256. http://www.nature.com/nature/journal/
190(1):73–94. https://doi.org/10.1016/0040-1951(91)90355-V v421/n6920/suppinfo/nature01351_S1.html
Cannat M, Sauter D, Mendel V, Ruellan E, Okino K, Escartin J, Com- Elkins L, Sims K, Prytulak J, Blichert-Toft J, Elliott T, Blusztajn J,
bier V, Baala M (2006) Modes of seafloor generation at a melt- Fretzdorff S, Reagan M, Haase K, Humphris S (2014) Melt gen-
poor ultraslow-spreading ridge. Geology 34(7):605–608 eration beneath Arctic ridges: implications from U decay series
Cherkaoui AS, Wilcock WS, Dunn RA, Toomey DR (2003) A numeri- disequilibria in the Mohns, Knipovich, and Gakkel Ridges. Geo-
cal model of hydrothermal cooling and crustal accretion at a chim Cosmochim Acta 127:140–170
fast spreading mid-ocean ridge. Geochem Geophys Geosyst Engdahl ER, van der Hilst R, Buland R (1998) Global teleseismic
4(9):8616 earthquake relocation with improved travel times and procedures
Coakley BJ, Cochran JR (1998) Gravity evidence of very thin crust for depth determination. Bull Seismol Soc Am 88(3):722–743
at the Gakkel Ridge (Arctic Ocean). Earth Planet Sci Lett Erdmann M, France L, Fischer LA, Deloule E, Koepke J (2017) Trace
162(1):81–95 elements in anatectic products at the roof of mid-ocean ridge
Coogan LA (2003) Contaminating the lower crust in the Oman ophi- magma chambers: an experimental study. Chem Geol 456:43–57.
olite. Geology 31(12):1065–1068. https://doi.org/10.1130/ https://doi.org/10.1016/j.chemgeo.2017.03.004
g20129.1 Expedition Scientific Party (2005) Oceanic core complex formation,
Coogan LA, Saunders AD, Kempton PD, Norry MJ (2000) Evidence Atlantis Massif, Mid-Atlantic Ridge: drilling into the footwall
from oceanic gabbros for porous melt migration within a crystal and hanging wall of a tectonic exposure of deep, young oceanic
mush beneath the Mid-Atlantic Ridge. Geochem Geophys Geo- lithosphere to study deformation, alteration, and melt genera-
syst 1(9):1044. https://doi.org/10.1029/2000GC000072 tion. IODP Prelim Rep. https://doi.org/10.2204/iodp.pr.305.2005
Coogan LA, Mitchell NC, O’Hara MJ (2003) Roof assimilation at Fietzke J, Frische M (2016) Experimental evaluation of elemental
fast spreading ridges: an investigation combining geophysical, behavior during LA-ICP-MS: influences of plasma conditions
geochemical, and field evidence. J Geophys Res Solid Earth and limits of plasma robustness. J Anal At Spectrom. 31(1):234–
108(B1):ECV2-1–ECV2-14 244. https://doi.org/10.1039/C5JA00253B
DeMets C, Gordon RG, Argus DF (2010) Geologically current plate Fietzke J, Liebetrau V, Günther D, Gürs K, Hametner K, Zumholz K,
motions. Geophys J Int 181(1):1–80 Hansteen T, Eisenhauer A (2008) An alternative data acquisition
Detrick RS, Mutter JC, Buhl P, Kim II (1990) No evidence from mul- and evaluation strategy for improved isotope ratio precision using
tichannel reflection data for a crustal magma chamber in the LA-MC-ICP-MS applied to stable and radiogenic strontium iso-
MARK area on the Mid-Atlantic Ridge. Nature 347(6288):61–64 topes in carbonates. J Anal At Spectrom 23(7):955–961
Devey CW, Lackschewitz KS, Baker E (2005) Hydrothermal and Fischer LA, Erdmann M, France L, Wolff PE, Deloule E, Zhang C,
volcanic activity found on the Southern Mid-Atlantic Ridge. Godard M, Koepke J (2016) Trace element evidence for anatexis
EOS Trans Am Geophys Union 86(22):209–212. https://doi. at oceanic magma chamber roofs and the role of partial melts
org/10.1029/2005EO220001 for contamination of fresh MORB. Lithos 260:1–8. https://doi.
Devey CW, German C, Haase K, Lackschewitz K, Melchert B, Con- org/10.1016/j.lithos.2016.05.001
nelly D (2010) The relationships between volcanism, tectonism, Fouquet Y (1997) Where are the large hydrothermal sulphide depos-
and hydrothermal activity on the southern equatorial Mid-Atlan- its in the oceans? Philos Trans R Soc Math Phys Eng Sci
tic Ridge. In: Rona PA, Devey CW, Dyment J, Murton BJ (eds) 355(1723):427–441
Diversity of hydrothermal systems on slow spreading ocean Fournier R (1987) Conceptual models of brine evolution in
ridges, pp 133–152 magmatic-hydrothermal systems. US Geol Surv Prof Pap
Dick HJB, Natland JH, Alt JC, Bach W, Bideau D, Gee JS, Haggas S, 1350(2):1487–1506
Hertogen JGH, Hirth G, Holm PM, Ildefonse B, Iturrino GJ, John France L, Ildefonse B, Koepke J (2009) Interactions between magma
BE, Kelley DS, Kikawa E, Kingdon A, LeRoux PJ, Maeda J, and hydrothermal system in Oman ophiolite and in IODP Hole

13
Contrib Mineral Petrol (2017) 172:97 Page 21 of 23  97

1256D: Fossilization of a dynamic melt lens at fast spread- Hoernle K, Hauff F, Kokfelt TF, Haase K, Garbe-Schönberg D, Werner
ing ridges. Geochem Geophys Geosyst 10(10).https://doi. R (2011) On- and off-axis chemical heterogeneities along the
org/10.1029/2009GC002652 South Atlantic Mid-Ocean-ridge (5–11°S): shallow or deep recy-
France L, Koepke J, Ildefonse B, Cichy SB, Deschamps F (2010) cling of ocean crust and/or intraplate volcanism? Earth Planet Sci
Hydrous partial melting in the sheeted dike complex at fast Lett 306(1–2):86–97. https://doi.org/10.1016/j.epsl.2011.03.032
spreading ridges: experimental and natural observations. Con- Hofmann AW (1988) Chemical differentiation of the Earth: the rela-
trib Miner Petrol 160(5):683–704 tionship between mantle, continental crust, and oceanic crust.
France L, Koepke J, MacLeod CJ, Ildefonse B, Godard M, Deloule E Earth Planet Sci Lett 90(3):297–314
(2014) Contamination of MORB by anatexis of magma cham- Hofmann AW, Jochum KP, Seufert M, White WM (1986) Nb
ber roof rocks: constraints from a geochemical study of experi- and Pb in oceanic basalts: new constraints on mantle evo-
mental melts and associated residues. Lithos 202–203:120– lution. Earth Planet Sci Lett 79(1–2):33–45. https://doi.
137. https://doi.org/10.1016/j.lithos.2014.05.018 org/10.1016/0012-821X(86)90038-5
German C, Connelly D, Evans A, Parson L (2002) Hydrothermal International Seismological Centre (2011) On-line bulletin. Interna-
activity on the southern Mid-Atlantic Ridge. In: AGU Fall tional Seismological Centre. http://www.isc.ac.uk. Accessed 19
Meeting Abstracts, vol 1. p 1361 Dec 2012
Gillis KM, Thompson G, Kelley DS (1993) A view of the lower Ito E, Harris DM, Anderson AT Jr (1983) Alteration of oce-
crustal component of hydrothermal systems at the Mid-Atlan- anic crust and geologic cycling of chlorine and water.
tic Ridge. J Geophys Res Solid Earth 98(B11):19597–19619. Geochim Cosmochim Acta 47(9):1613–1624. https://doi.
https://doi.org/10.1029/93JB01717 org/10.1016/0016-7037(83)90188-6
Gillis KM, Coogan LA, Chaussidon M (2003) Volatile element (B, Jarosewich E, Nelen JA, Norberg JA (1980) Reference samples for elec-
Cl, F) behaviour in the roof of an axial magma chamber from tron microprobe analysis. Geostand Newsl 4(1):43–47. https://
the East Pacific Rise. Earth Planet Sci Lett 213(3–4):447–462. doi.org/10.1111/j.1751-908X.1980.tb00273.x
https://doi.org/10.1016/s0012-821x(03)00346-7 Jenner FE, O’Neill HSC (2012) Analysis of 60 elements in 616 ocean
Godard M, Awaji S, Hansen H, Hellebrand E, Brunelli D, Johnson K, floor basaltic glasses. Geochem Geophys Geosyst 13(2):Q02005
Yamasaki T, Maeda J, Abratis M, Christie D, Kato Y, Mariet Jochum KP, Stoll B, Herwig K, Willbold M, Hofmann AW, Amini M,
C, Rosner M (2009) Geochemistry of a long in situ section of Aarburg S, Abouchami W, Hellebrand E, Mocek B (2006) MPI-
intrusive slow-spread oceanic lithosphere: results from IODP DING reference glasses for in situ microanalysis: new reference
Site U1309 (Atlantis Massif, 30°N Mid-Atlantic-ridge). Earth values for element concentrations and isotope ratios. Geochem
Planet Sci Lett 279(1–2):110–122. https://doi.org/10.1016/j. Geophys Geosyst 7(2). https://doi.org/10.1029/2005GC001060
epsl.2008.12.034 Jokat W, Ritzmann O, Schmidt-Aursch MC, Drachev S, Gauger S,
Goldstein SL, Soffer G, Langmuir CH, Lehnert KA, Graham Snow J (2003) Geophysical evidence for reduced melt produc-
DW, Michael PJ (2008) Origin of a `Southern Hemisphere’ tion on the Arctic ultraslow Gakkel mid-ocean ridge. Nature
geochemical signature in the Arctic upper mantle. Nature 423(6943):962–965. http://www.nature.com/nature/journal/
453(7191):89–93. http://www.nature.com/nature/journal/v453/ v423/n6943/suppinfo/nature01706_S1.html
n7191/suppinfo/nature06919_S1.html Kawahata H, Nohara M, Ishizuka H, Hasebe S, Chiba H (2001) Sr
Gregory RT, Taylor HP (1981) An oxygen isotope profile in a section isotope geochemistry and hydrothermal alteration of the Oman
of Cretaceous oceanic crust, Samail Ophiolite, Oman: evidence ophiolite. J Geophys Res Solid Earth 106(B6):11083–11099.
for δ18O buffering of the oceans by deep (> 5 km) seawater- https://doi.org/10.1029/2000JB900456
hydrothermal circulation at mid-ocean ridges. J Geophys Res Kelemen PB, Aharonov E (1998) Periodic formation of magma frac-
Solid Earth 86(B4):2737–2755 tures and generation of layered gabbros in the lower crust beneath
Grevemeyer I, Reston TJ, Moeller S (2013) Microseismicity of the oceanic spreading ridges. In: Buck WR, Delaney PT, Karson
Mid-Atlantic Ridge at 7°S–8°15′S and at the Logatchev Massif JA, Lagabrielle Y (eds) Faulting and magmatism at Mid-Ocean
oceanic core complex at 14°40′N–14°50′N. Geochem Geophys ridges. American Geophysical Union, pp 267–289
Geosyst 14(9):3532–3554 Kelemen PB, Shimizu N, Salters VJM (1995) Extraction of mid-ocean-
Grimes CB, John BE, Cheadle MJ, Wooden JL (2008) Protracted ridge basalt from the upwelling mantle by focused flow of melt
construction of gabbroic crust at a slow spreading ridge: con- in dunite channels. Nature 375(6534):747–753
straints from 206Pb/238U zircon ages from Atlantis Massif and Kelemen PB, Koga K, Shimizu N (1997) Geochemistry of gabbro
IODP Hole U1309D (30 N, MAR). Geochem Geophys Geosyst sills in the crust-mantle transition zone of the Oman ophiolite:
9(8). https://doi.org/10.1029/2008GC002063 implications for the origin of the oceanic lower crust. Earth
Haase K, Brandl PA, Devey CW, Hauff F, Melchert B, Garbe-Schön- Planet Sci Lett 146(3–4):475–488. https://doi.org/10.1016/
berg D, Kokfelt T, Paulick H (2016) Compositional variation S0012-821X(96)00235-X
and 226Ra-230Th model ages of axial lavas from the south- Kendrick MA, Arculus R, Burnard P, Honda M (2013) Quantify-
ern Mid-Atlantic Ridge, 8°48′S. Geochem Geophys Geosyst ing brine assimilation by submarine magmas: examples from
17(1):199–218 the Galápagos Spreading Centre and Lau Basin. Geochim
Harper GD (1985) Tectonics of slow spreading mid-ocean ridges Cosmochim Acta 123:150–165. https://doi.org/10.1016/j.
and consequences of a variable depth to the brittle/ductile gca.2013.09.012
transition. Tectonics 4(4):395–409. https://doi.org/10.1029/ Kendrick MA, Hemond C, Kamenetsky VS, Danyushevsky L, Devey
TC004i004p00395 CW, Rodemann T, Jackson MG, Perfit MR (2017) Seawater
Hart S, Erlank A, Kable E (1974) Sea floor basalt alteration: cycled throughout Earth’s mantle in partially serpentinized
some chemical and Sr isotopic effects. Contrib Miner Petrol lithosphere. Nat Geosci 10:222–228. https://doi.org/10.1038/
44(3):219–230 NGEO2902
Hasenclever J, Theissen-Krah S, Rüpke LH, Morgan JP, Iyer K, Kent AJR, Norman MD, Hutcheon ID, Stolper EM (1999) Assimilation
Petersen S, Devey CW (2014) Hybrid on-axis plus ridge-per- of seawater-derived components in an oceanic volcano: evidence
pendicular circulation reconciles hydrothermal flow observa- from matrix glasses and glass inclusions from Loihi seamount,
tions at fast spreading ridges. Nature 508:508–512 Hawaii. Chem Geol 156(1–4):299–319. https://doi.org/10.1016/
S0009-2541(98)00188-0

13
97   Page 22 of 23 Contrib Mineral Petrol (2017) 172:97

Klein EM, Langmuir CH (1987) Global correlations of ocean ridge Michael P (1995) Regionally distinctive sources of depleted
basalt chemistry with axial depth and crustal thickness. J Geo- MORB: evidence from trace elements and H2O.
phys Res Solid Earth 92(B8):8089–8115. https://doi.org/10.1029/ Earth Planet Sci Lett 131(3–4):301–320. https://doi.
JB092iB08p08089 org/10.1016/0012-821X(95)00023-6
Klügel A (1998) Reactions between mantle xenoliths and host Michael PJ, Cornell WC (1998) Influence of spreading rate and
magma beneath La Palma (Canary Islands): constraints on magma supply on crystallization and assimilation beneath
magma ascent rates and crustal reservoirs. Contrib Miner Petrol mid-ocean ridges: evidence from chlorine and major ele-
131(2–3):237–257 ment chemistry of mid-ocean ridge basalts. J Geophys Res
Kovalenko VI, Naumov VB, Girnis AV, Dorofeeva VA, Yarmolyuk 103(B8):18325–18356. https://doi.org/10.1029/98jb00791
VV (2006) Estimation of the average contents of ­H2O, Cl, F, Michael PJ, Graham DW (2015) The behavior and concentration of
and S in the depleted mantle on the basis of the compositions ­CO 2 in the suboceanic mantle: inferences from undegassed
of melt inclusions and quenched glasses of mid-ocean ridge ocean ridge and ocean island basalts. Lithos 236–237:338–351.
basalts. Geochem Int 44(3):209–231. https://doi.org/10.1134/ https://doi.org/10.1016/j.lithos.2015.08.020
s0016702906030013 Morgan JP, Chen YJ (1993) The genesis of oceanic crust: Magma
Kvassnes AS, Grove T (2008) How partial melts of mafic lower crust injection, hydrothermal circulation, and crustal flow. J Geophys
affect ascending magmas at oceanic ridges. Contrib Miner Petrol Res Solid Earth 98(B4):6283–6297
156(1):49–71. https://doi.org/10.1007/s00410-007-0273-x Michael PJ, Schilling J-G (1989) Chlorine in mid-ocean ridge mag-
Labidi J, Cartigny P, Hamelin C, Moreira M, Dosso L (2014) Sulfur mas: evidence for assimilation of seawater-influenced compo-
isotope budget (32S, 33S, 34S and 36S) in Pacific-Antarctic ridge nents. Geochim Cosmochim Acta 53(12):3131–3143. https://
basalts: a record of mantle source heterogeneity and hydrother- doi.org/10.1016/0016-7037(89)90094-x
mal sulfide assimilation. Geochim Cosmochim Acta 133:47–67. Michael P, Langmuir C, Dick H, Snow J, Goldstein S, Graham D,
https://doi.org/10.1016/j.gca.2014.02.023 Lehnert K, Kurras G, Jokat W, Mühe R (2003) Magmatic and
le Roux PJ, Shirey SB, Hauri EH, Perfit MR, Bender JF (2006) The amagmatic seafloor generation at the ultraslow-spreading Gak-
effects of variable sources, processes and contaminants on the kel Ridge, Arctic Ocean. Nature 423(6943):956–961
composition of northern EPR MORB (8–10°N and 12–14°N): Minshull T, Bruguier N, Brozena J (1998) ridge-plume interactions
evidence from volatiles (­ H2O, ­CO2, S) and halogens (F, Cl). Earth or mantle heterogeneity near Ascension Island? Geology
Planet Sci Lett 251(3–4):209–231. https://doi.org/10.1016/j. 26(2):115–118
epsl.2006.09.012 Wise SA, Watters RL (2012) Certificate of Analysis, Standard Refer-
Lecuyer C, Reynard B (1996) High-temperature alteration of oceanic ence Material 610. National Institute of Standards and Tech-
gabbros by seawater (Hess Deep, Ocean Drilling Program Leg nology. http://www.nist.gov/srm
147): evidence from oxygen isotopes and elemental fluxes. J Möller H (2002) Magma Genesis and Mantle Source at the Mid-
Geophys Res Solid Earth 101(B7):15883–15897. https://doi. Atlantic Ridge East of Ascension Island. Dissertation at Chris-
org/10.1029/96JB00950 tian-Albrechts-Univeristät zu Kiel
Lehnert K, Su Y, Langmuir C, Sarbas B, Nohl U (2000) A global Montési LG, Behn MD (2007) Mantle flow and melting underneath
geochemical database structure for rocks. Geochem Geophys oblique and ultraslow Mid-Ocean ridges. Geophys Res Lett
Geosyst 1(5):1012. https://doi.org/10.1029/1999GC000026 34(24). https://doi.org/10.1029/2007GL031067
Lissenberg CJ, Dick HJB (2008) Melt–rock reaction in the lower Mottl M (2003) Partitioning of energy and mass fluxes between
oceanic crust and its implications for the genesis of mid-ocean mid-ocean ridge axes and flanks at high and low temperature.
ridge basalt. Earth Planet Sci Lett 271(1–4):311–325. https://doi. In: Halbach P, Tunnicliffe V, Hein JR (eds) Energy and mass
org/10.1016/j.epsl.2008.04.023 transfer in marine hydrothermal systems. Dahlem University
Lissenberg CJ, Bédard JH, van Staal CR (2004) The structure and geo- Press, Berlin, pp 271–286
chemistry of the gabbro zone of the Annieopsquotch ophiolite, Nehlig P, Juteau T (1988) Flow porosities, permeabilities and
newfoundland: implications for lower crustal accretion at spread- preliminary data on fluid inclusions and fossil thermal
ing ridges. Earth Planet Sci Lett 229(1–2):105–123. https://doi. gradients in the crustal sequence of the Sumail ophiolite
org/10.1016/j.epsl.2004.10.029 (Oman). Tectonophysics 151(1–4):199–221. https://doi.
Lissenberg CJ, Rioux M, Shimizu N, Bowring SA, Mével C (2009) Zir- org/10.1016/0040-1951(88)90246-6
con dating of oceanic crustal accretion. Science 323(5917):1048– Nicolas A, Mainprice D, Boudier F (2003) High-temperature sea-
1050. https://doi.org/10.1126/science.1167330 water circulation throughout crust of oceanic ridges: a model
Maclennan J, Hulme T, Singh SC (2005) Cooling of the lower oceanic derived from the Oman ophiolites. J Geophys Res Solid Earth
crust. Geology 33(5):357–366. https://doi.org/10.1130/g21207.1 108(B8):2371. https://doi.org/10.1029/2002JB002094
McDonough WF, Sun SS (1995) The composition of the Niu Y, Hekinian R (1997) Spreading-rate dependence of the extent
Ear th. Chem Geol 120(3–4):223–253. https://doi. of mantle melting beneath ocean ridges. Nature 385:326–329
org/10.1016/0009-2541(94)00140-4 Palme H, O’Neill HSC (2003) Cosmochemical estimates of mantle
Melchert B, Devey CW, German C, Lackschewitz K, Seifert R, Wal- composition. In: Heinrich DH, Karl KT (eds) Treatise on geo-
ter M, Mertens C, Yoerger D, Baker E, Paulick H (2008) First chemistry, vol 2. Pergamon, Oxford, pp 1–38
evidence for high-temperature off-axis venting of deep crustal/ Paulick H, Münker C, Schuth S (2010) The influence of small-scale
mantle heat: the Nibelungen hydrothermal field, southern Mid- mantle heterogeneities on Mid-Ocean Ridge volcanism: evi-
Atlantic Ridge. Earth Planet Sci Lett 275(1):61–69 dence from the southern Mid-Atlantic Ridge (7 30′S to 11 30′S)
Mével C, Cannat M (1991) Lithospheric stretching and hydrothermal and Ascension Island. Earth Planet Sci Lett 296(3–4):299–310.
processes in oceanic gabbros from slow-spreading ridges. In: https://doi.org/10.1016/j.epsl.2010.05.009
Peters TJ, Nicolas A, Coleman R (eds) Ophiolite genesis and Pontbriand CW, Soule SA, Sohn RA, Humphris SE, Kunz C, Singh
evolution of the oceanic lithosphere. Springer, pp 293–312 H, Nakamura Ki, Jakobsson M, Shank T (2012) Effusive
Meyer P, Dick HB, Thompson G (1989) Cumulate gabbros from the and explosive volcanism on the ultraslow-spreading Gakkel
Southwest Indian ridge, 54°S–7°16′E: implications for mag- Ridge, 85°E. Geochem Geophys Geosyst 13(10). https://doi.
matic processes at a slow spreading ridge. Contrib Miner Petrol org/10.1029/2012GC004187
103(1):44–63. https://doi.org/10.1007/BF00371364

13
Contrib Mineral Petrol (2017) 172:97 Page 23 of 23  97

Putirka KD (2008) Thermometers and barometers for volcanic systems. during emplacement. Earth Planet Sci Lett 252(3–4):289–307.
Rev Miner Geochem 69(1):61–120 https://doi.org/10.1016/j.epsl.2006.09.043
Putirka K, Johnson M, Kinzler R, Longhi J, Walker D (1996) Thermo- Stakes D, Vanko DA (1986) Multistage hydrothermal altera-
barometry of mafic igneous rocks based on clinopyroxene-liquid tion of gabbroic rocks from the failed Mathematician
equilibria, 0–30 kbar. Contrib Miner Petrol 123(1):92–108 ridge. Earth Planet Sci Lett 79(1–2):75–92. https://doi.
Reid I, Jackson H (1981) Oceanic spreading rate and crustal thickness. org/10.1016/0012-821X(86)90042-7
Mar Geophys Res 5(2):165–172 Stein CA, Stein S (1994) Constraints on hydrothermal heat flux through
Ridley WI, Perfit MR, Smith MC, Fornari DJ (2006) Magmatic pro- the oceanic lithosphere from global heat flow. J Geophys Res
cesses in developing oceanic crust revealed in a cumulate xeno- Solid Earth 99(B2):3081–3095
lith collected at the East Pacific Rise, 9°50′N. Geochem Geophys Stolper E (1982) The speciation of water in silicate melts. Geochim
Geosyst 7(12):Q12O04. https://doi.org/10.1029/2006GC001316 Cosmochim Acta 46(12):2609–2620
Roeder P, Emslie R (1970) Olivine-liquid equilibrium. Contrib Miner Stroncik NA, Haase KM (2004) Chlorine in oceanic intraplate basalts:
Petrol 29(4):275–289 constraints on mantle sources and recycling processes. Geology
Rubin KH, Sinton JM (2007) Inferences on mid-ocean ridge ther- 32(11):945–948. https://doi.org/10.1130/g21027.1
mal and magmatic structure from MORB compositions. Earth Stroncik NA, Niedermann S (2016) Atmospheric contamination of the
Planet Sci Lett 260(1–2):257–276. https://doi.org/10.1016/j. primary Ne and Ar signal in mid-ocean ridge basalts and its
epsl.2007.05.035 implications for ocean crust formation. Geochim Cosmochim
Rutherford MJ (2008) Magma ascent rates. Rev Miner Geochem Acta 172:306–321. https://doi.org/10.1016/j.gca.2015.09.016
69(1):241–271 Sun WD, Binns RA, Fan AC, Kamenetsky VS, Wysoczanski R, Wei
Ryabchikov ID (2001) Deep geospheres and ore genesis. Geol Rudn GJ, Hu YH, Arculus RJ (2007) Chlorine in submarine volcanic
Mestorozhd 43:195–207 glasses from the eastern manus basin. Geochim Cosmochim Acta
Saal AE, Hauri EH, Langmuir CH, Perfit MR (2002) Vapour under 71(6):1542–1552. https://doi.org/10.1016/j.gca.2006.12.003
saturation in primitive mid-ocean-ridge basalt and the volatile Urann BM, Le Roux V, Hammond K, Marschall HR, Lee C-TA, Mon-
content of Earth’s upper mantle. Nature 419(6906):451–455. teleone BD (2017) Fluorine and chlorine in mantle minerals and
http://www.nature.com/nature/journal/v419/n6906/suppinfo/ the halogen budget of the Earth’s mantle. Contrib Miner Petrol
nature01073_S1.html 172(7):51. https://doi.org/10.1007/s00410-017-1368-7
Salters VJM, Stracke A (2004) Composition of the depleted man- van der Zwan FM, Fietzke J, Devey CW (2012) Precise measurement of
tle. Geochem Geophys Geosyst 5(5):Q05B07 https://doi. low (< 100 ppm) chlorine concentrations in submarine basaltic
org/10.1029/2003gc000597 glass by electron microprobe. J Anal At Spectrom 27:1966–1974
Sanfilippo A, Tribuzio R, Tiepolo M (2014) Mantle–crust interactions van der Zwan FM, Devey CW, Augustin N, Almeev RR, Bantan RA,
in the oceanic lithosphere: constraints from minor and trace Basaham A (2015) Hydrothermal activity at the ultraslow- to
elements in olivine. Geochim Cosmochim Acta 141:423–439. slow-spreading Red Sea Rift traced by chlorine in basalt. Chem
https://doi.org/10.1016/j.gca.2014.06.012 Geol 405:63–81. https://doi.org/10.1016/j.chemgeo.2015.04.001
Sano T, Miyoshi M, Ingle S, Banerjee NR, Ishimoto M, Fukuoka Vine F, Moores E (1972) A model for the gross structure, petrology,
T (2008) Boron and chlorine contents of upper oceanic crust: and magnetic properties of oceanic crust. Geol Soc Am Mem
Basement samples from IODP Hole 1256D. Geochem Geophys 132:195–206
Geosyst 9(12):Q12O15. https://doi.org/10.1029/2008GC002182 Wanless V, Perfit M, Ridley W, Klein E (2010) Dacite petrogenesis
Sauter D, Cannat M, Rouméjon S, Andreani M, Birot D, Bronner A, on mid-ocean ridges: evidence for oceanic crustal melting and
Brunelli D, Carlut J, Delacour A, Guyader V (2013) Continu- assimilation. J Petrol 51(12):2377–2410
ous exhumation of mantle-derived rocks at the Southwest Indian Wanless V, Perfit M, Ridley W, Wallace P, Grimes C, Klein E (2011)
Ridge for 11 million years. Nat Geosci 6(4):314 Volatile abundances and oxygen isotopes in basaltic to dacitic
Schramm B, Devey CW, Gillis KM, Lackschewitz K (2005) Quanti- lavas on mid-ocean ridges: the role of assimilation at spreading
tative assessment of chemical and mineralogical changes due centers. Chem Geol 287(1):54–65
to progressive low-temperature alteration of East Pacific Rise Weaver SJ, Langmuir CH (1990) Calculation of phase equilibrium in
basalts from 0 to 9 Ma. Chem Geol 218(3–4):281–313. https:// mineral-melt systems. Comput Geosci 16(1):1–19
doi.org/10.1016/j.chemgeo.2005.01.011 Yamashita S, Kitamura T, Kusakabe M (1997) Infrared spectroscopy
Shaw C, Klügel A (2002) The pressure and temperature conditions of hydrous glasses of arc magma compositions. Geochem J Jpn
and timing of glass formation in mantle-derived xenoliths from 31:169–174
Baarley, West Eifel, Germany: the case for amphibole break- Zhang C, Wang L-X, Marks MAW, France L, Koepke J (2017) Vola-
down, lava infiltration and mineral-melt reaction. Miner Petrol tiles ­(CO2, S, F, Cl, Br) in the dike-gabbro transition zone at
74(2–4):163–187 IODP Hole 1256D: magmatic imprint versus hydrothermal influ-
Shaw AM, Behn MD, Humphris SE, Sohn RA, Gregg PM (2010) Deep ence at fast-spreading mid-ocean ridge. Chem Geol 459:43–60.
pooling of low degree melts and volatile fluxes at the 85 E seg- https://doi.org/10.1016/j.chemgeo.2017.04.002
ment of the Gakkel Ridge: evidence from olivine-hosted melt Zhao M, Canales JP, Sohn RA (2012) Three-dimensional seis-
inclusions and glasses. Earth Planet Sci Lett 289(3):311–322 mic structure of a Mid-Atlantic Ridge segment characterized
Shishkina T, Botcharnikov R, Holtz F, Almeev R, Portnyagin MV by active detachment faulting (Trans-Atlantic Geotraverse,
(2010) Solubility of ­H2O- and ­CO2-bearing fluids in tholeiitic 25°55′N–26°20′N). Geochem Geophys Geosyst 13(11). https://
basalts at pressures up to 500 MPa. Chem Geol 277(1):115–125 doi.org/10.1029/2012GC004454
Soule SA, Fornari DJ, Perfit MR, Ridley WI, Reed MH, Cann JR
(2006) Incorporation of seawater into mid-ocean ridge lava flows

13

You might also like