Role of Metasomatism in The Development of The East African Rift at The Northern Tanzanian Divergence: Insights From 3D Magnetotelluric Modeling
Role of Metasomatism in The Development of The East African Rift at The Northern Tanzanian Divergence: Insights From 3D Magnetotelluric Modeling
10.1029/2023GC011191
                                                African Rift at the Northern Tanzanian Divergence: Insights
Key Points:                                     From 3D Magnetotelluric Modeling
• 3D magnetotelluric models of North
  Tanzanian Divergence are converted to         Sinan Özaydın1 , Kate Selway2,3, Stephen F. Foley4                , Isra S. Ezad4, William L. Griffin4,
  water in mantle models to map
                                                Pascal S. Tarits5, and Sophie Hautot6
  metasomatism in the region
• Melting events in the Mozambique              1
                                                 School of Geosciences, University of Sydney, Sydney, NSW, Australia, 2School of Natural Sciences, University of
  Belt caused metasomes to be destroyed
  and the lithospheric mantle to be             Tasmania, Hobart, TAS, Australia, 3Vox Geophysics, Perth, WA, Australia, 4School of Natural Sciences, Macquarie
  dehydrated                                    University, Sydney, NSW, Australia, 5Laboratoire Géosciences Océan, Institut Universitaire Européen de la Mer, Plouzané,
• The rifting in the region might be            France, 6IMAGIR sarl, Saint Renan, France
  limited if there is no supply of
  metasomatic material toward the rift
  zone                                          Abstract       The Northern Tanzanian Divergence in the East Africa Rift is arguably the best place on Earth to
                                                study the controls on rifting of thick lithosphere. Here, where the East Africa Rift intersects the Tanzanian
Supporting Information:                         Craton and the Mozambique Belt, the relationships between volcanism, faulting, pre‐existing structures and
Supporting Information may be found in          lithospheric thickness and composition can be observed. In this work, we carry out the first lithospheric‐scale
the online version of this article.
                                                3D magnetotelluric modeling of the Northern Tanzanian Divergence and combine the results with experimental
                                                electrical conductivity and petrology models to calculate mantle composition, which is also inferred in the
Correspondence to:                              craton from reanalysis of garnet xenocryst data. Our results show that metasomatic materials exist in the cratonic
S. Özaydın,
[email protected]
                                                lithospheric mantle and the relatively undeveloped southern part of the rift zone. However, the lithospheric
                                                mantle of the Mozambique Belt and the more developed northern section of the rift is more resistive and does
                                                not contain metasomatic phases. Combined with geochemical data from erupted lavas, these results suggest that,
Citation:
                                                in zones that have experienced voluminous Cenozoic magmatism, melting events have destroyed the metasomes
Özaydın, S., Selway, K., Foley, S. F.,
Ezad, I. S., Griffin, W. L., Tarits, P. S., &   and dehydrated the mantle. Since the presence of magma is a primary control of lithospheric strength, rifting
Hautot, S. (2024). Role of metasomatism         may become limited as the lithospheric mantle becomes dehydrated and harder to melt.
in the development of the East African Rift
at the Northern Tanzanian Divergence:
Insights from 3D magnetotelluric
                                                Plain Language Summary                    The motion of tectonic plates relies on a specific set of physical
modeling. Geochemistry, Geophysics,             conditions. Continental breakup or rifting occurs when certain parts of the lithosphere are weak, and when stress
Geosystems, 25, e2023GC011191. https://         applied to these regions is sufficient. Weaknesses in the lithosphere rely on its composition and pre‐existing
doi.org/10.1029/2023GC011191
                                                structures. We can image and analyze these features using the magnetotelluric method, a geophysical technique
Received 21 AUG 2023                            that maps electrical conductivity variations within the Earth. Our results show that compositionally weakening
Accepted 4 JAN 2024                             agents (metasomes) play an essential role in the development of the rift by making the mantle easier to melt. We
                                                also image some portions of the rift that do not contain such agents, suggesting that melts may have dried out
                                                these parts of the lithosphere, leaving a dry and resistive residue. This situation may indicate that melting in the
                                                region might be limited in the long run due to the absence of these materials.
                                                1. Introduction
                                                The initiation and evolution of intracontinental rifts are fundamental to the theory of plate tectonics. Most simply
                                                put, they begin to develop when the stresses applied can overcome the strength of the lithosphere. The intricate
                                                interplay between the distribution of stress, magmatism, pre‐existing lithospheric architecture, mantle compo-
                                                sition and rheology can result in a variety of rifting styles (Brune et al., 2023). The East African Rift System
                                                (EARS) is the largest active continental rift system in the world and displays various stages of rift development
                                                along its strike length (Boone et al., 2019). In this study, we focus on the region where the EARS meets the
© 2024 The Authors. Geochemistry,               Tanzanian Craton at the North Tanzanian Divergence (NTD). Here, the rift structure widens and there is an
Geophysics, Geosystems published by
Wiley Periodicals LLC on behalf of              increase in the geochemical variety and volume of volcanism. The geochemical characteristics of lavas and
American Geophysical Union.                     mantle xenoliths in the NTD reflect the evolving nature of the lithosphere through the Cenozoic in response to
This is an open access article under the        plate reorganization and plume impingement (Baptiste et al., 2015; Foley et al., 2012; Mana et al., 2015;
terms of the Creative Commons
Attribution License, which permits use,         Rooney, 2020).
distribution and reproduction in any
medium, provided the original work is           Magnetotellurics (MT) is a powerful tool to reveal the composition and architecture of the lithosphere‐
properly cited.                                 asthenosphere system (Naif et al., 2021; Selway, 2014; Selway & O’Donnell, 2019). It is especially sensitive
ÖZAYDIN ET AL.                                                                                                                                                     1 of 19
                                                                                                                                              15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                    10.1029/2023GC011191
                  to interconnected secondary conductive phases (e.g., melts, hydrous minerals) and water (OH bound to
                  nominally anhydrous minerals), which are often products of metasomatism. Since deformation by diffusion creep
                  in olivine is water‐dependent (Hirth & Kohlstedf, 2003), these metasomatized and hydrated regions are more
                  likely to be rheologically weaker. The same regions are also often more prone to melting since they contain
                  components such as H2O or CO2, which cause a substantial drop in the solidus temperature (Foley &
                  Pintér, 2018), and metasomatic phases that are easier to melt, such as hydrous pyroxenites (Foley et al., 2022).
                  Moreover, regions rich in metasomes are now envisaged as one of the reasons that such thick lithosphere can
                  initiate rifting (Foley & Fischer, 2017; Rooney, 2020), where the metasomes provide a weaker lithosphere either
                  through the existence of melt (Buck, 2006) or the combined effect of the hydrolytic weakening of olivine and/or
                  grain size reduction (Selway, 2015).
                  It has been demonstrated that MT can be useful in understanding the relationship between mantle composition and
                  magmatic processes. For instance, (Özaydın & Selway, 2022) used MT models to show that kimberlites might
                  exploit the “lithospheric fuel” frozen in metasomatized mantle in order to ascend. However, the effects of large‐
                  volume magmatism (e.g., basalts) on the composition of the lithosphere need to be better understood (Özaydın
                  et al., 2022), since they could possibly exhaust these metasomes and dehydrate the mantle as well. Consequences
                  of such dynamics may be crucial for rift development (Foley & Fischer, 2017; Muirhead et al., 2020).
                  Previous research in the NTD has combined MT and seismic tomography studies to reveal the existence of melt
                  within the crust of the rift zone, mainly around the Manyara fault (Clutier et al., 2021; Plasman et al., 2019; Reiss
                  et al., 2022; Tiberi et al., 2019). 2D modeling of long‐period MT data has also imaged the large‐scale lithospheric
                  structure in the area, with results suggesting that water content is higher in the cratonic lithosphere than the rift and
                  is not therefore the primary control on deformation localization (Selway, 2015).
                  Here, we image the deep electrical structure in NTD with 3D MT modeling utilizing the combined MT data sets of
                  Selway (2015) and Plasman et al. (2019). Electrical conductivity variations in the mantle can be used to make
                  quantified interpretations of composition employing experimental electrical conductivity and petrology studies
                  (Özaydın & Selway, 2020; Selway, 2014). We made these calculations using the software MATE (Özaydın &
                  Selway, 2020) and a geophysically‐constrained thermal model (Afonso et al., 2022). In the cratonic domain, we
                  also constructed lithological sections from the garnet xenocryst database (Griffin et al., 1991; O’Reilly &
                  Griffin, 1996), using the methods described in Griffin et al. (2002). Using the MT model, we checked for the
                  presence of water, melt, and other conductive phases. We compared these results with the current knowledge from
                  geochronological studies, geochemical modeling, mantle xenoliths/xenocrysts and other geophysical studies to
                  form a better understanding of the geodynamics of the region.
                  2. Geological Background
                  The rocks of north‐eastern Tanzania record more than 2.5 billion years of continental evolution, including cra-
                  tonization, rifting, collision and multiple episodes of reactivation of lithospheric‐scale structures. The Tanzania
                  Craton amalgamated by c. 2.6 Ga (Chesley et al., 1999; Manya et al., 2006; Thomas et al., 2016) and geophysical
                  studies suggest that the lithosphere has seismic wavespeeds typical of cratonic domains to depths of at least
                  150 km (e.g., Afonso et al., 2022; Emry et al., 2019; Mulibo & Nyblade, 2013b; O’Donnell et al., 2013), with low
                  surface heat flows (Nyblade, 1997). The cratonic lithosphere has been sampled by Jurassic to Quaternary
                  kimberlite magmatism and Tertiary to Recent rift‐related volcanism (e.g., Foley et al., 2012; Rooney, 2020).
                  Petrographic and geochemical analyses of xenoliths show evidence that the cratonic lithosphere has been met-
                  asomatized during multiple events since the Archean (e.g., Aulbach et al., 2011; Baptiste et al., 2015; Chesley
                  et al., 1999; Koornneef et al., 2009; Stachel et al., 1998). This agrees with MT models that show the cratonic
                  lithosphere is likely to be hydrated and metasomatized (Selway, 2014, 2015).
                  After cratonization, the first major tectonic event to affect the north‐east Tanzania Craton was the Neoproterozoic
                  East African (or Pan‐African) Orogen, which was associated with the amalgamation of East and West Gondwana
                  and formed the Mozambique Belt as part of an extensive band of deformed lithosphere that extends from East
                  Africa into Antarctica (e.g., Grantham et al., 2003; Stern, 1994). Despite the Neoproterozoic timing of defor-
                  mation, isotopic and geochronological data from across the Mozambique Belt show that it consists largely of
                  reworked Archean lithosphere, including protoliths with ages similar to Tanzanian Craton rocks (e.g.,
                  Maboko, 2000; Thomas et al., 2016), implying that the Tanzania Craton may have originally extended further to
ÖZAYDIN ET AL.                                                                                                                     2 of 19
                                                                                                                                          15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                 10.1029/2023GC011191
                  the east than its present extent. Peak East African Orogen metamorphism occurred at c. 640 Ma (Muhongo
                  et al., 2001) and was followed by a period of relative quiescence.
                  North‐eastern Tanzania is currently being deformed as part of the East African Rift (Ebinger, 2012), the most
                  extensive and best exposed active continental rift on Earth, which extends from Ethiopia to Malawi. Seismic
                  tomography models show relatively slow wavespeeds at lower to upper mantle depths beneath central and eastern
                  Africa and a thinned mantle transition zone, which is interpreted to be caused by a hot mantle plume impinging on
                  the base of the African lithosphere (e.g., Emry et al., 2019; Hansen et al., 2012; Mulibo & Nyblade, 2013a;
                  O’Donnell et al., 2013; Ritsema et al., 1999). The geochemistry of Cenozoic magmas support the existence of a
                  plume underlying East Africa, with evidence for elevated mantle temperatures and plume magma sources
                  (Rooney, 2020; Rooney et al., 2012). The initial plume impact is interpreted to have occurred c. 30–40 Myr ago
                  (Ebinger & Sleep, 1998; Hofmann et al., 1997). Geodynamic models suggest that the present rifting is dominantly
                  caused by deviatoric stresses induced by plume‐related uplift (Koptev et al., 2016; Stamps et al., 2014) and much
                  of the deformation has reactivated pre‐existing structures, including those formed during the East African Orogen,
                  suggesting that they have continued to be zones of lithospheric weakness (e.g., Daly et al., 1989; Tommasi &
                  Vauchez, 2001).
                  The character of the East African Rift changes markedly along its extent, from incipient oceanic spreading in the
                  northern part of the rift to the first gasps of magmatism in the Rungwe Province, south of the Tanzanian Craton.
                  Where it meets the Tanzanian Craton, the rift bifurcates into Eastern and Western Branches, seemingly following
                  weaker lithosphere that surrounds the strong craton. The NTD is the section of the Eastern Branch in north‐eastern
                  Tanzania and is characterized by a relatively broad zone of volcanism and block faulting (Clutier et al., 2021; Le
                  Gall et al., 2008; Tiberi et al., 2019). The Eyasi and Manyara rifts extend into the eastern margin of the Tanzania
                  Craton, and the volcanic centers at Labait and Hanang have sampled cratonic lithosphere that is being actively
                  impacted by the plume (Le Gall et al., 2008), making this an ideal location to study the controls on continental
                  rifting. As is the case for the broader rift, faulting and volcanism tend to follow pre‐existing zones of weakness.
                  Volcanism in the NTD initiated at c. 6 Ma at locations in the west of the NTD close to the edge of the craton; with
                  time new volcanic centers have erupted further to the east, while volcanism has continued and spread in the west,
                  including to Oldoinyo Lengai, the only active carbonatitic volcano on Earth (Mana et al., 2015). While the timing
                  of faulting is harder to quantify, most faulting appears to have occurred in the last 4 Myr and movement on
                  individual faults appears to be temporally correlated with volcanism (Le Gall et al., 2008).
                  3. Methods
                  3.1. Magnetotelluric Data and Modeling
                  MT data used in this study are a compilation from stations published previously (Plasman et al., 2019; Sel-
                  way, 2015). These data include the 21 long‐period stations reported in Selway (2015), which were from a regional
                  study around the North Tanzanian Divergence (green triangles, Figure 1b), and 24 broad‐band stations from
                  Plasman et al. (2019), from a study focused more on the rift zone (white triangles, Figure 1b). The long‐period
                  data also included tipper data, which was included in the inversion. Some of the stations are excluded from the
                  broad‐band data set due to the limitations that arise from cell sizes constructed for the large‐scale model. For this
                  problem, we eliminated the most locally similar‐looking stations to avoid loss of information in the long periods
                  (Figure 2).
                  Three‐dimensional magnetotelluric inversions were carried out using the ModEM algorithm (Kelbert
                  et al., 2014). The 3D nature of the area is apparent both from the high phase tensor skew degrees observed at all
                  frequency ranges and highly variable strike directions (Figure 3), which require three‐dimensional MT modeling
                  techniques to obtain reliable results (Booker, 2014). The horizontal discretization in the core region was chosen to
                  be 5 km in both directions. Outside the core region, we inserted seven padding cells with widths that increase by a
                  factor of 1.5. In the vertical domain, we used a total of 53 cells with depths that increase by a factor of 1.15,
                  starting from the third cell with a thickness of 150 m. We choose to add three 50 m layer cells at the top of the
                  model to partially mitigate the problems that may arose from galvanic distortion (Kelbert et al., 2014). The total
                  depth in the vertical direction is ∼705 km. The ocean is dealt with by inserting fixed 0.3 Ωm valued cells ac-
                  cording to global topography/bathymetry data ETOPO1 (Amante & Eakins, 2009).
ÖZAYDIN ET AL.                                                                                                                 3 of 19
                                                                                                                                                                        15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                      Geochemistry, Geophysics, Geosystems                                                          10.1029/2023GC011191
Figure 1. Maps showing the study area: (a) Main tectonic units, active volcanoes and Cenozoic volcanic units of Eastern African Rift System overlain on LAB depths
derived from the study of Afonso et al. (2022). (b) Geological map of the main study area alongside MT stations, kimberlite locations (Giuliani & Pearson, 2019), and
Cenozoic volcanic rocks derived from the GeoRoc database (Lehnert et al., 2000). MKF: Mwadui Kimberlite Field, EKF: Eyasi Kimberlite Field.
                                          We performed the inversion in two main stages. In the first stage, we inverted only the tippers from the long‐
                                          period stations at 25 frequencies between 1 and 10,000 s. The errors on the tipper were fixed at a single value
                                          of 0.05. The initial model was constructed as a homogeneous half‐space with 210 Ωm resistivity, which is the
                                          median determinant apparent resistivity value calculated from impedance tensors from all stations between 100
                                          and 10,000 s. An isotropic covariance matrix was constructed using the value of 0.5 in all directions. The initial
                                          regularization parameter (λ) was selected as 10 and set to decrease by a factor of 5 when the RMS difference
                                          between subsequent iterations became less than 0.002. The inversion ran for 55 iterations until it converged to a
                                          final RMS value of 1.59, starting from 3.48.
                                          In the second stage of inversion, we inverted both the full impedance tensor and tipper data, using the result from
                                                                                                                             √̅̅̅̅̅̅̅
                                          the tipper inversion as the initial model. The error floors were chosen to be 5% of Z xy Z yx for all impedance tensor
                                          elements. We chose to decrease the constraints on tipper data with an error of 0.15 to give more weight to im-
                                          pedances in this stage of the inversion. Twenty‐five frequencies between 1 and 10,000 s were inverted. The
                                          inversion was conducted with the same regularization parameter and reduction scheme as the first stage of the
                                          inversion, and the isotropic covariance value was reduced to 0.3. We also ran the inversions with covariance
                                          values and observed no crucial differences that would affect our interpretations (Figures S2 and S3 in Supporting
                                          Information S1). Inversion started with an initial RMS value of 16.25 and finalized with a value of 2.52. The local
                                          RMS values can be seen in the RMS map of impedance fittings (Figure S6 in Supporting Information S1) and
                                          individual graphs (Figures S53–S115 in Supporting Information S1).
                                          Sensitivity tests on the resistivity models were performed on the conductors C1 and CM to test their robustness. We
                                          masked these conductive regions with blocks in electrical resistivities varying between 1 and 10,000 Ωm (Figures
                                          S4 and S5 in Supporting Information S1). The results demonstrated that both conductors appear robust and
                                          applicable for interpretation in the models.
ÖZAYDIN ET AL.                                                                                                                                                4 of 19
                                                                                                                                                                           15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                       Geochemistry, Geophysics, Geosystems                                                           10.1029/2023GC011191
Figure 2. Apparent resistivity and phase curves of selected off‐diagonal stations used in the inversions. More information on data can be found in Figures S7–S52 of the
Supporting Information S1. LP: Long Period, BB: Broad‐band.
ÖZAYDIN ET AL.                                                                                                                                                   5 of 19
                                                                                                                                                        15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                            10.1029/2023GC011191
                  Figure 3. Phase tensor ellipses filled with skew angles (β°) for (a) 10, (b) 100, (c) 1,000, and (d) 2,500 s. Absolute skew angles
                  above the value of 3° are considered to be out of the practical limits of reliable 2D MT modeling (Booker, 2014).
                                                                                        ΔH
                                                                        σ = σ0 exp(        )                                                      (1)
                                                                                        RT
                  where σ0 is pre‐exponent (S/m), ΔH is the activation enthalpy, and R is the gas constant. For the NAMs, the
                  electrical conduction processes can be described as a summation of three conduction mechanisms that operate on
                  different temperature levels: ionic (σion), polaron (σion) and proton (σpro) conduction (Equation 2).
                  For the depths in which we are interested in this study, polaron and proton conduction are the most relevant
                  conduction mechanisms, and relate to the electrical conductivity of dry and hydrated minerals, respectively.
                  While electrical conductivities of silicate minerals have high temperature dependencies (high activation
                  enthalpy), there are some other phases such as graphite and sulphides that have very low temperature de-
                  pendencies (low activation enthalpy; Özaydın & Selway, 2020).
                  The model for water distribution among NAMs is constructed using the water partitioning coefficients shown in
                  Table 1. We sought solutions of water contents between a dry lithosphere and bulk water solubility values
                  calculated using water partitioning coefficients, based upon the sub‐solidus olivine water solubility model of
                  Padrón‐Navarta and Hermann (2017). Since this model limits water solubility to low levels (several tens of ppm)
                  in the shallow lithospheric mantle, we cannot use water content as a proxy for metasomatism in most of the
                  uppermost lithosphere (<70–90 km). Therefore, one has to be mindful while interpreting our figures in terms of
                  how metasomatism translates to different signatures going from the lower lithosphere (>70–90 km) to above. One
                  can most easily do this by looking at resistivity and water content maps in tandem and checking whether a
                  conductor originates from water‐rich/metasomatized areas.
                   Table 1
                   Mineral Water Partitioning Coefficients Used in This Study
                                                                   Water partitioning coefficient                                 Reference
                   Orthopyroxene/Olivine                                    Dopx∕ol = 5.6                                  Demouchy et al. (2017)
                   Clinopyroxene/Olivine                               Dcpx∕ol = Dopx∕ol × 1.9                             Demouchy et al. (2017)
                   Garnet/Olivine                                            Dgt/ol = 0.8                                   Novella et al. (2014)
ÖZAYDIN ET AL.                                                                                                                                6 of 19
                                                                                                                                             15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                    10.1029/2023GC011191
                  We have used the electrical conductivity models of Gardés et al. (2014), Dai and Karato (2009a), Liu et al. (2019),
                  and Dai and Karato (2009b) for olivine, orthopyroxene, clinopyroxene and garnet, respectively. For the con-
                  ductivity of phlogopite, the model of Li et al. (2017) was utilized with a fluorine content of 0.52 w.t. % (average
                  fluorine content value in mantle rocks, Özaydın et al., 2022). All electrical conductivity models are corrected for
                  the water measurement calibrations of Withers et al. (2012) for olivine, and Bell et al. (1995) for pyroxenes and
                  garnet if needed. A lherzolitic composition was used to calculate the water contents in the region (Table S1 in
                  Supporting Information S1). The use of different mantle peridotitic modal compositions has been shown to have a
                  negligible effect on understanding the variations of metasomatism (Özaydın & Selway, 2022). The Generalized
                  Archie's Law (Glover, 2010) was used for phase mixing to calculate bulk electrical conductivity. Interconnections
                  for the minerals are constructed with cementation components of m = 2 for orthopyroxene, m = 4 for clino-
                  pyroxene and garnet and m < 1 for olivine, which gives results close to Hashin‐Shtrikman lower‐bound
                  (Özaydın & Selway, 2020). The thermal model used in these calculations is taken from multi‐observable
                  probabilistic inversions of Afonso et al. (2022).
                  Water measurements made on xenoliths from Labait, Lashaine, Olmani, and Pello Hill were also compared with
                  the MT‐derived water models (Baptiste et al., 2015; Hui et al., 2015). To compare the bulk water contents, we
                  used only the water measurements made on orthopyroxenes since they represent a more reliable water recorder
                  than olivine (Yang et al., 2019). We converted orthopyroxene water contents to bulk water contents assuming the
                  same partition coefficients used in MT‐derived water models and composition. We have also used individual
                  sample‐based modal compositions in this conversion, which are indicated with different symbols (Figure 6). The
                  orthopyroxene water contents reported in Baptiste et al. (2015) were measured with calibration of Pater-
                  son (1982), which is known to underestimate the water contents (Demouchy & Bolfan‐Casanova, 2016). We
                  corrected these values by multiplying the water contents by 3 (Demouchy & Bolfan‐Casanova, 2016).
                  The constructed geotherm, alongside garnet major‐ and trace‐element compositions, can be used to make
                  compositional sections of the lithosphere. We use two garnet xenocryst classification schemes to understand the
                  nature of the lithosphere. The first one is the CARP method (Cluster Analysis by Regressive Partitioning; Griffin
                  et al., 2002), which shows the proportion of garnets derived from five different lithologies: (a) Depleted harz-
                  burgites, (b) Depleted lherzolites, (c) Depleted lherzolites with phlogopite metasomatism, (d) Fertile lherzolites,
                  and (e) Melt‐metasomatized lithologies. The other classification scheme (Figures 10c and 10d), taken from the
                  work of Grütter et al. (2004), determines whether the host lithology is harzburgite, lherzolite, wehrlite, mega-
                  crystic, Ti‐metasomatized, pyroxenite, or eclogite based on garnet CaO Cr2O3 contents. Thermobarometry and
                  classifications were made using the python library Thermobar (Wieser et al., 2022), specifically using the Cr‐
                  pyrope garnet thermobarometry method of Ryan et al. (1996).
ÖZAYDIN ET AL.                                                                                                                    7 of 19
                                                                                                                                                                     15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                      Geochemistry, Geophysics, Geosystems                                                        10.1029/2023GC011191
Figure 4. Mantle composition sections derived from kimberlite‐derived garnet xenocrysts on the Tanzanian Craton (Green stars, Figure 1), alongside geophysical and
xenolith‐based geotherms. (a) CARP mantle composition section using the methodology of Griffin et al. (2002). (b) The number of samples analyzed in with CARP
methodology. (c) CaO Cr2O3 based mantle section based on the classification scheme of Grütter et al. (2004). (d) Geotherms constructed with garnet‐xenocrysts
alongside geotherms extracted from the geophysical thermal model of Afonso et al. (2022), for the Tanzanian Craton (gray) and Mozambique Belt (red). The pink
diamond markers indicate the pyroxene‐based thermobarometric calculations derived from xenoliths from Quaternary lavas in Labait (Lee & Rudnick, 1999). The black
dashed line indicates the base of the depleted lithosphere.
                                         shown in Figure 5. The first two electrical resistivity slices, down to 25 km, include information from earthquake
                                         epicenters in the region, while the deeper slices include information on the distribution of solidified magmatic
                                         products. We also show vertical slices from the electrical resistivity model where MT stations are denser and form
                                         a profile (Figures 6 and 7).
ÖZAYDIN ET AL.                                                                                                                                             8 of 19
                                                                                                                                                                 15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                    Geochemistry, Geophysics, Geosystems                                                     10.1029/2023GC011191
                                                                            Along the A‐A′ transect, we observe a good correlation with our calculated
                                                                            water content variations, the LAB depths acquired from the study of Afonso
                                                                            et al. (2022) and the base of the depleted lithosphere from garnet xenocrysts
                                                                            (Figure 4). Water content calculations demonstrate that the cratonic litho-
                                                                            spheric mantle beneath this region is variably metasomatized (e.g., near C1).
                                                                            Since hydrous minerals can also enhance mantle conductivity, we also
                                                                            considered combinations of phlogopite and water in NAMs that could
                                                                            explain the conductivity structures along this transect. To do this, we first
                                                                            added phlogopite with different degrees of interconnectivity to a dry lher-
                                                                            zolite matrix and compared the resulting conductivities with those observed
                                                                            at 10 km above the LAB of Afonso et al. (2022) (Figure 6e). If the
                                                                            phlogopite grains are very well connected (Archie's Law m = 1.1) very low
                                                                            volume percentages throughout the transect can match the observed con-
                                                                            ductivities (<0.5%). On the other hand, non‐connected phlogopites (m = 5)
                                                                            require unrealistically large volumetric abundances to explain the conduc-
                                                                            tivities (>15%). We calculated the effect of 15% phlogopite in a lherzolite
                                                                            matrix on seismic velocities with the toolbox of Abers and Hacker (2016):
                                                                            this showed that such high phlogopite contents would result in a VS of
                                                                            ∼4.17 km/s beneath C1 at 140 km depth, which is much lower than what
                                                                            was observed (∼4.5–4.6 km/s, O’Donnell et al., 2013). The results, overall,
                                                                            suggest that amounts of phlogopite that fit the electrical conductivity values
                                                                            and the seismic model can only be present if they are moderately inter-
                                                                            connected (m = 2.5).
ÖZAYDIN ET AL.                                                                                                                                        9 of 19
                                                                                                                                                                                 15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                        Geochemistry, Geophysics, Geosystems                                                               10.1029/2023GC011191
Figure 6. MT slice along the A‐A′ profile and other estimated and cataloged properties. (a) Histogram showing the number of kimberlites within the 50 km proximity of
the profile. Since there are no kimberlites within the Eastern portion, the profile is cut after Fault Eyasi. (b) Elevation along the profile. (c) MT slice along the profile,
also showing LAB and MOHO acquired from Afonso et al. (2022). (d) Water content estimation along the profile, where the shaded regions with white indicate the
places below the water solubility limit. Melted lithosphere portions at different stages (Stage 1–4) envisioned by the study of Mana et al. (2015) are indicated on the slice.
Bulk water solubility limits calculated with Padrón‐Navarta and Hermann (2017) for end‐member Craton and Mozambique Belt geotherms are also shown on the right‐
hand side of the figure. (e) Estimated volumetric abundances of phlogopite (% 0.52 F) to fit the MT model 10 km below the LAB with different interconnection
(m) values.
                                            most electrically conductive/heavily metasomatized regions such as the region around the lithosphere‐spanning
                                            conductor C1.
ÖZAYDIN ET AL.                                                                                                                                                       10 of 19
                                                                                                                                                                     15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                       Geochemistry, Geophysics, Geosystems                                                      10.1029/2023GC011191
                                                                                 The conductor C2 sits under the Northern Crater Highlands where volcanoes
                                                                                 Oldoinyo Lengai, Embagai, Loolmasin, and Olmati are situated, while C3 sits
                                                                                 under the volcanoes Essimingor, Tarosero, and Monduli. These crustal con-
                                                                                 ductors may represent the presence of active magma or the effects of past
                                                                                 magmatism in the form of crystalline conductive material. The metasomat-
                                                                                 ized portion of the deeper lithospheric mantle (CM) would then represent a
                                                                                 plausible place for incipient melts to form in response to an oxidized
                                                                                 (CO2 + H2O) solidus (Pintér et al., 2021) since our calculations fitting
                                                                                 electrical conductivities with water‐induced melting do not reduce the solidus
                                                                                 enough to meet the local geotherm and induce melting (Figure 8). Another
Figure 7. MT and estimated bulk water content slices along (a) C‐C′, (b) B‐
B′, and (c) D‐D′ profiles. Melted regions envisioned by the study of             possibility could be that magmas form by the melting of hydrous pyroxenites
Baudouin and Parat (2020) for Kwaraha and Labait Volcanoes are indicated         in the region, which matches with the geotherm at CM at these depths
with red water content slice of (a). (d) Modeled water content profiles around   (Figure 8; Foley et al., 2022). Following this, we can suggest that these melts
the Labait Volcano compared to xenolith water contents Hui et al. (2015).        originated around CM in response to a sub‐lithospheric heat source and may
We used the orthopyroxene water measurements to convert them to bulk
                                                                                 have used oblique lithospheric weakness zones represented by CM 2 and
water contents with composition used in water models (blue circles) and
individual compositions for each sample indicated in the study.                  CM 3 to ascend, forming conductors by mineralizing conductive phases along
                                                                                 the way (Figure 9).
                                           Alternatively, we can assume that the magmas do not necessarily travel along these oblique lithospheric zones of
                                           weakness and are instead emplaced vertically upwards from their source regions. This would be likely to happen
                                           if the lithosphere beneath the volcanoes is also metasomatized/hydrated to account for metasome‐induced
                                           incipient melting. While we observe hydrated zones beneath most of the volcanoes in the region, the litho-
                                           sphere beneath the Northern Crater Highlands, where Oldoinyo Lengai, Embagan, and Kerimasi are situated, as
                                           well as the lithosphere underlying Mt Meru and Mt Kilimanjaro, appears less hydrous (Figures 5–7) One potential
                                           pattern, which would require more extensive MT coverage for confirmation, is that dehydrated mantle appears to
                                           coincide with the more recent volcanism, while more hydrated mantle coincides with dormant volcanism in the
                                           southwest of the model region (Figures 6 and 7; Mana et al., 2015). In this model, recent melt generation would
                                           have increased the resistivity of these regions by melting metasomatic phases and partitioning water into the melt
                                           (e.g., Novella et al., 2014) and melts would have migrated without precipitating any conductive phases. If this
                                           were to be true, we can envisage that the lithospheric pathways may be likely to be more conductive if they use a
                                           pre‐existing weakness zone such as CM 2 and CM 3, occupying a plate boundary (Manyara Fault, Figure 9). In
                                           either case, it is possible that narrow conductive zones caused by magma/fluid infiltration may exist but are too
ÖZAYDIN ET AL.                                                                                                                                          11 of 19
                                                                                                                                                                   15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                       Geochemistry, Geophysics, Geosystems                                                    10.1029/2023GC011191
ÖZAYDIN ET AL.                                                                                                                                        12 of 19
                                                                                                                                                                        15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                      Geochemistry, Geophysics, Geosystems                                                          10.1029/2023GC011191
Figure 9. 3D representation of the MT model in the eastern portion of the area (see shaded area in the inset map) structured with (a) 3D MT model slices, and (b) the
interpretive sketch dependent on the results of this study.
                                          destroyed and may induce melting in the region as for the younger ages of the lavas of Labait, Hanang, and
                                          Kwaraha (Mana et al., 2015).
                                          4.4. Implications for the Stability of the Tanzanian Craton and Rift Propagation
                                          Tectonically, one of the most striking features about northeast Tanzania is the vastly different responses to stress
                                          between the Tanzanian Craton and the adjacent lithosphere in the Mozambique Belt and Rift Basin. The
                                          Mozambique Belt underwent deformation of the whole lithosphere during the Pan‐African Orogeny and is
                                          currently being deformed again during rifting. In contrast, the Tanzanian Craton, or at least the surviving portion
                                          of the Tanzanian Craton (Ebinger et al., 1997), has remained relatively tectonically stable during these events,
                                          seemingly able to withstand stresses that elsewhere are enough to produce continental‐scale deformation. The
                                          apparent strength of the Tanzanian Craton is highlighted by the fact that the East Africa Rift, after extending
                                          approximately linearly from the Main Ethiopian Rift to the northern margin of the Tanzanian Craton, then bi-
                                          furcates around the Tanzanian Craton and splits into the Eastern Branch, which we image here, and the Western
                                          Branch (Figure 1). This contrasting behavior indicates that there are significant rheological differences between
                                          the Tanzanian Craton and the adjacent lithosphere.
                                          The dominant conclusion from numerous studies investigating the rheological response to rifting in this region
                                          is not that it is unusual that the Tanzanian Craton is stable, but rather that it is instead surprising that the
                                          adjacent lithosphere of the Eastern and Western Branches is rifting (e.g., Behn et al., 2006; Buck, 2006; Koptev
                                          et al., 2016; O’Donnell et al., 2016). The stresses in the region are tensional and primarily derived from
                                          gradients in the gravitational potential associated with the uplift of the East African Plateau (Craig et al., 2011;
                                          Rajaonarison et al., 2021; Stamps et al., 2014). These stresses provide only a fraction of the stress theoretically
                                          needed to rift thick (>100 km thickness) and melt‐free lithosphere. Adding to this challenge, as pointed out by
                                          Selway et al. (2014) and Selway (2015) and confirmed in the results shown here, is that much of the litho-
                                          spheric mantle beneath the Eastern Branch is dehydrated. Since water content in olivine is one of the major
ÖZAYDIN ET AL.                                                                                                                                               13 of 19
                                                                                                                                              15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                    10.1029/2023GC011191
                  controls on mantle rheology (e.g., Hirth & Kohlstedf, 2003), higher water contents in the cratonic mantle
                  should, all else being equal, make it weaker than the Eastern Branch lithospheric mantle. Selway (2015)
                  suggested that small grain sizes in the Eastern Branch lithosphere, which may still remain as a scar from Pan‐
                  African deformation, could contribute to weakening the Eastern Branch lithosphere and could outweigh the
                  impact of water content (e.g., Ramirez et al., 2022). Mantle xenoliths from the volcanoes along the rift zone
                  (Pello Hill and Elodoi) do indeed dominantly have porphyroclastic textures with smaller olivine grain sizes
                  (0.4–2 mm Baptiste et al., 2015). In the less developed parts of the rift zone toward the south (Labait Volcano),
                  porphyroclastic textures only exist in the samples with the highest equilibrium temperatures (1450C°), whereas
                  larger grain sizes (5–15 mm) were observed in lithospheric samples (Vauchez et al., 2005). Xenolith samples
                  from Lashaine, east of the rift valley, show similar textures to lithospheric samples from Labait (Baptiste
                  et al., 2015). Even though they occur in close proximity in space and time to Lashaine, Olmani xenoliths consist
                  of porphyroclastic to coarse textures with small grain sizes, suggesting that the area east of the rift valley
                  experienced heterogeneous distribution of stress and, therefore, that localized deformation might have occurred
                  (Baptiste et al., 2015).
                  Although fine grain sizes in the rift zone will reduce mantle viscosity, most calculations suggest that the East
                  African lithosphere should still be too strong to rift in the presence of the available stresses (e.g., O’Donnell
                  et al., 2016). Instead, the presence of melt is the most effective mechanism to reduce the strength of the Tanzanian
                  lithosphere enough that it deforms (e.g., Buck, 2006; O’Donnell et al., 2016; Reiss et al., 2022). Correlations
                  between the timing of volcanism and faulting in the NTD (e.g., Le Gall et al., 2008) lend support to the idea that
                  the East African Rift lithosphere is weakened by active magmatism. The MT results presented here can improve
                  our understanding of the distribution of melt within the lithospheric mantle of the NTD and of the likely
                  development of large‐scale deformation. Within the cratonic mantle lithosphere, although resistivities are low, the
                  lack of high‐volume Cenozoic surface volcanism and the high seismic velocities suggest that there is no melt
                  present and that the low resistivities are attributable to solid‐state causes. Within the Eastern Branch, despite the
                  generally higher resistivities, the recent and active volcanism (Mana et al., 2015), low seismic velocities (Clutier
                  et al., 2021; Tiberi et al., 2019), and high seismic attenuation (Reiss et al., 2022) suggest that melt is present in the
                  lithosphere. The seismic and MT observations can be reconciled if melt is present in a sufficiently small volume
                  that it resides in triple junctions but does not coat grain boundaries (e.g., Selway & O’Donnell, 2019) and the
                  mantle conductor CM (Figure 6) may indicate regions with slightly higher and more interconnected melt
                  concentrations.
                  Geochemical investigations of magmatism in the East Africa Rift suggest that thermal input has occurred in
                  pulses (e.g., Rooney, 2020). During later thermal pulses, portions of the lithospheric mantle that were meta-
                  somatized in earlier pulses are progressively melted until, eventually, the lithospheric mantle is depleted and
                  magmas are sourced from the sub‐lithospheric mantle (e.g., Mana et al., 2015; Rooney, 2020). The relatively
                  dehydrated nature of the lithospheric mantle in the Mozambique Belt suggests that this process is already well‐
                  developed and might support the numerical models of Koptev et al. (2016), where plume impingement causes a
                  large pond of magma initiating the rifting process but also dehydrating and destroying the metasomes in the Belt.
                  Magma currently within the lithospheric mantle may contain significant sub‐lithospheric material (Muirhead
                  et al., 2020) and may be scavenging any remaining metasomatized material from the lithosphere (Mana
                  et al., 2015). This dehydration of the lithosphere suggests that, unless there are large future thermal pulses to inject
                  significant volumes of sub‐lithospheric melt into the Eastern Branch lithospheric mantle, future volcanism in this
                  part of the Eastern Branch may be limited. Given that magma is a primary control on lithospheric strength, this
                  may mean that the extent of rifting in the Eastern Branch will also be limited. The outlook for rifting and
                  deformation would differ if the plume were to migrate toward parts of the lithospheric mantle that still contain
                  metasomatic phases, including the adjacent cratonic lithosphere, inducing redox melting (Foley & Fischer, 2017;
                  Muirhead et al., 2020). The effect of grain sizes, hydrolytic weakening of olivine, structural inheritance and
                  existence of melt should be tested with geodynamic modeling using the results of this study to further constrain
                  the mechanisms accommodating rifting in this region.
                  5. Conclusions
                  Newly developed 3D MT models of the Northern Tanzanian Divergence are subjected to quantified in-
                  terpretations. The most important conclusions of this study are:
ÖZAYDIN ET AL.                                                                                                                   14 of 19
                                                                                                                                                                                                      15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                                               Geochemistry, Geophysics, Geosystems                                                                       10.1029/2023GC011191
                                                •   The cratonic region displays signs of a metasomatized lithosphere throughout, especially around the promi-
                                                    nent conductor C1. The large conductor C1 lies above the water solubility limit of NAMs, suggesting other
                                                    minor conductive phases, particularly graphite and sulphides, might be responsible for its formation.
                                                •   Kimberlites on the craton appear on the metasomatized/hydrated portions of the lithosphere and avoid the
                                                    prominent conductor C1, which aligns with the previously observed relationship between electrical conduc-
                                                    tivity and kimberlite distribution around the world (Özaydın & Selway, 2022).
                                                •   The garnet xenocryst section demonstrates a layered lithosphere where the mantle below ∼160 km is abun-
                                                    dantly melt‐metasomatized, which matches mantle xenolith observations from the region (e.g., Lee & Rud-
                                                    nick, 1999). This boundary also coincides with geophysical interpretations of the current LAB and displays a
                                                    transition from a hydrated/metasomatized to a dry lithosphere.
                                                •   The rift zone is electrically heterogeneous, where primary conductors follow the fault zones. In the deeper rift
                                                    zone, the conductor CM appears, which correlates well with the suggested melt‐bearing regions from seismic
                                                    tomography studies (Clutier et al., 2021; Reiss et al., 2022). As well as melt, the conductive nature of the
                                                    region may indicate the presence of metasomes that are likely to reduce the solidus and induce melting in the
                                                    presence of a thermal anomaly.
                                                •   The mantle beneath the northern rift zone (proximal to Oldoinyo Lengai) and the region spanning Mt Essi-
                                                    mingor to Mt Kilimanjaro appears to be dehydrated. This might indicate that melting in these areas caused
                                                    metasomes to be destroyed, and consequently the mantle to be dehydrated. The fact that mantle xenolith water
                                                    contents also do not match the MT‐derived water contents in these regions might indicate that the area has
                                                    evolved significantly in composition since eruption of the xenoliths.
                                                •   The dehydrated nature of the Mozambique Belt suggests that rifting might be accommodated through melting
                                                    of metasomes across the rift zone. In the absence of an enhanced thermal anomaly, the rifting in the Eastern
                                                    Branch might be limited in future geological times.
Acknowledgments                                 References
This study was financially supported by
the Covid Recovery Fellowship scheme            Abers, G. A., & Hacker, B. R. (2016). A MATLAB toolbox and Excel workbook for calculating the densities, seismic wave speeds, and major
run by Macquarie University. All original         element composition of minerals and rocks at pressure and temperature. Geochemistry, Geophysics, Geosystems, 17(2), 616–624. https://doi.
MT data collection was supported by the           org/10.1002/2015gc006171
government of Tanzania through the              Acocella, V. (2014). Structural control on magmatism along divergent and convergent plate boundaries: Overview, model, problems. Earth‐
Tanzanian Commission for Science and              Science Reviews, 136, 226–288. https://doi.org/10.1016/j.earscirev.2014.05.006
Technology (COSTECH) and the                    Afonso, J. C., Ben‐Mansour, W., O’Reilly, S. Y., Griffin, W. L., Salajeghegh, F., Foley, S. F., et al. (2022). Thermochemical structure and
Geological Survey of Tanzania. We are             evolution of cratonic lithosphere in central and southern Africa. Nature Geoscience, 15(5), 405–410. https://doi.org/10.1038/s41561‐022‐
grateful for the generous assistance of the       00929‐y
University of Dar es Salaam, local schools,     Amante, C., & Eakins, B. W. (2009). ETOPO1 1 arc‐minute global relief model: Procedures, data sources and analysis. NOAA technical
residents and land‐owners, and particularly       memorandum NESDIS NGDC‐24 (Vol. 10, p. V5C8276M). National Geophysical Data Center, NOAA.
Mr Mtelela Khalfan (UDSM) whose                 Aulbach, S., Lin, A.‐B., Weiss, Y., & Yaxley, G. (2020). Wehrlites from continental mantle monitor the passage and degassing of carbonated
extraordinary skills, patience, good nature       melts. Geochemical Perspective Letters, 30–34. https://doi.org/10.7185/geochemlet.2031
and hard work enabled these data sets to be     Aulbach, S., Rudnick, R. L., & McDonough, W. F. (2011). Evolution of the lithospheric mantle beneath the East African Rift in Tanzania and its
collected. We thank Rob Evans for their           potential signatures in rift magmas. Geological Society of America Special Paper, 478, 105–125.
constructive review of this article.            Baptiste, V., Tommasi, A., Vauchez, A., Demouchy, S., & Rudnick, R. L. (2015). Deformation, hydration, and anisotropy of the lithospheric
Open access publishing facilitated by The         mantle in an active rift: Constraints from mantle xenoliths from the North Tanzanian divergence of the East African Rift. Tectonophysics, 639,
University of Sydney, as part of the Wiley ‐      34–55. https://doi.org/10.1016/j.tecto.2014.11.011
The University of Sydney agreement via          Baudouin, C., & Parat, F. (2020). Phlogopite‐olivine nephelinites erupted during early stage rifting, North Tanzanian Divergence. Frontiers in
the Council of Australian University              Earth Science, 8, 277. https://doi.org/10.3389/feart.2020.00277
Librarians.                                     Baudouin, C., Parat, F., Denis, C. M., & Mangasini, F. (2016). Nephelinite lavas at early stage of rift initiation (Hanang Volcano, North Tanzanian
                                                  Divergence). Contributions to Mineralogy and Petrology, 171(7), 1–20. https://doi.org/10.1007/s00410‐016‐1273‐5
                                                Behn, M. D., Buck, W. R., & Sacks, I. S. (2006). Topographic controls on dike injection in volcanic rift zones. Earth and Planetary Science
                                                  Letters, 246(3–4), 188–196. https://doi.org/10.1016/j.epsl.2006.04.005
                                                Bell, D. R., Ihinger, P. D., & Rossman, G. R. (1995). Quantitative analysis of trace OH in garnet and pyroxenes. American Mineralogist, 80(5–6),
                                                  465–474. https://doi.org/10.2138/am‐1995‐5‐607
ÖZAYDIN ET AL.                                                                                                                                                                         15 of 19
                                                                                                                                                                          15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                                         10.1029/2023GC011191
                  Booker, J. R. (2014). The magnetotelluric phase tensor: A critical review. Surveys in Geophysics, 35(1), 7–40. https://doi.org/10.1007/s10712‐
                    013‐9234‐2
                  Boone, S. C., Kohn, B. P., Gleadow, A. J., Morley, C. K., Seiler, C., & Foster, D. A. (2019). Birth of the East African Rift system: Nucleation of
                    magmatism and strain in the Turkana Depression. Geology, 47(9), 886–890. https://doi.org/10.1130/g46468.1
                  Brune, S., Kolawole, F., Olive, J.‐A., Stamps, D. S., Buck, W. R., Buiter, S. J. H., et al. (2023). Geodynamics of continental rift initiation and
                    evolution. Nature Reviews Earth & Environment, 4(4), 235–253. https://doi.org/10.1038/s43017‐023‐00391‐3
                  Buck, W. (2006). The role of magma in the development of the Afro‐Arabian Rift System. Geological Society, London, Special Publications,
                    259(1), 43–54. https://doi.org/10.1144/gsl.sp.2006.259.01.05
                  Chesley, J. T., Rudnick, R. L., & Lee, C.‐T. (1999). Re‐Os systematics of mantle xenoliths from the East African Rift: Age, structure, and history
                    of the Tanzanian craton. Geochimica et Cosmochimica Acta, 63(7–8), 1203–1217. https://doi.org/10.1016/s0016‐7037(99)00004‐6
                  Clutier, A., Gautier, S., & Tiberi, C. (2021). Hybrid local and teleseismic P‐wave tomography in North Tanzania: Role of inherited structures and
                    magmatism on continental rifting. Geophysical Journal International, 224(3), 1588–1606. https://doi.org/10.1093/gji/ggaa538
                  Craig, T., Jackson, J., Priestley, K., & McKenzie, D. (2011). Earthquake distribution patterns in Africa: Their relationship to variations in lith-
                    ospheric and geological structure, and their rheological implications. Geophysical Journal International, 185(1), 403–434. https://doi.org/10.
                    1111/j.1365‐246x.2011.04950.x
                  Dai, L., Hu, H., Jiang, J., Sun, W., Li, H., Wang, M., et al. (2020). An overview of the experimental studies on the electrical conductivity of major
                    minerals in the upper mantle and transition zone. Materials, 13(2), 408. https://doi.org/10.3390/ma13020408
                  Dai, L., Hu, H., Sun, W., Li, H., Liu, C., & Wang, M. (2019). Influence of high conductive magnetite impurity on the electrical conductivity of dry
                    olivine aggregates at high temperature and high pressure. Minerals, 9(1), 44. https://doi.org/10.3390/min9010044
                  Dai, L., & Karato, S.‐I. (2009a). Electrical conductivity of orthopyroxene: Implications for the water content of the asthenosphere. Proceedings of
                    the Japan Academy, Series B, 85(10), 466–475. https://doi.org/10.2183/pjab.85.466
                  Dai, L., & Karato, S.‐I. (2009b). Electrical conductivity of pyrope‐rich garnet at high temperature and high pressure. Physics of the Earth and
                    Planetary Interiors, 176(1–2), 83–88. https://doi.org/10.1016/j.pepi.2009.04.002
                  Daly, M., Chorowicz, J., & Fairhead, J. (1989). Rift basin evolution in Africa: The influence of reactivated steep basement shear zones. Geological
                    Society, London, Special Publications, 44(1), 309–334. https://doi.org/10.1144/gsl.sp.1989.044.01.17
                  Dambly, M., Samrock, F., Grayver, A., & Saar, M. O. (2023). Insights on the interplay of rifting, transcrustal magmatism and formation of
                    geothermal resources in the central segment of the Ethiopian Rift revealed by 3‐D magnetotelluric imaging. Journal of Geophysical Research:
                    Solid Earth, 128(7), e2022JB025849. https://doi.org/10.1029/2022jb025849
                  Dasgupta, R., Mallik, A., Tsuno, K., Withers, A. C., Hirth, G., & Hirschmann, M. M. (2013). Carbon‐dioxide‐rich silicate melt in the Earth’s upper
                    mantle. Nature, 493(7431), 211–215. https://doi.org/10.1038/nature11731
                  Demouchy, S., & Alard, O. (2021). Hydrogen, trace, and ultra‐trace element distribution in natural olivines. Contributions to Mineralogy and
                    Petrology, 176(4), 26. https://doi.org/10.1007/s00410‐021‐01778‐5
                  Demouchy, S., & Bolfan‐Casanova, N. (2016). Distribution and transport of hydrogen in the lithospheric mantle: A review. Lithos, 240, 402–425.
                    https://doi.org/10.1016/j.lithos.2015.11.012
                  Demouchy, S., Shcheka, S., Denis, C. M., & Thoraval, C. (2017). Subsolidus hydrogen partitioning between nominally anhydrous minerals in
                    garnet‐bearing peridotite. American Mineralogist, 102(9), 1822–1831. https://doi.org/10.2138/am‐2017‐6089
                  Denis, C. M., Demouchy, S., & Alard, O. (2018). Heterogeneous hydrogen distribution in orthopyroxene from veined mantle peridotite (San
                    Carlos, Arizona): Impact of melt‐rock interactions. Lithos, 302, 298–311. https://doi.org/10.1016/j.lithos.2018.01.007
                  Ebinger, C. (2012). Evolution of the Cenozoic East African Rift system: Cratons, plumes and continental breakup. In Regional geology and
                    tectonics: Phanerozoic rift systems and sedimentary basins (Vol. 1, pp. 132–162). Elsevier.
                  Ebinger, C., Djomani, Y. P., Mbede, E., Foster, A., & Dawson, J. (1997). Rifting Archaean lithosphere: The Eyasi‐Manyara‐Natron rifts, East
                    Africa. Journal of the Geological Society, 154(6), 947–960. https://doi.org/10.1144/gsjgs.154.6.0947
                  Ebinger, C., & Sleep, N. (1998). Cenozoic magmatism throughout East Africa resulting from impact of a single plume. Nature, 395(6704),
                    788–791. https://doi.org/10.1038/27417
                  Emry, E. L., Shen, Y., Nyblade, A. A., Flinders, A., & Bao, X. (2019). Upper mantle Earth structure in Africa from full‐wave ambient noise
                    tomography. Geochemistry, Geophysics, Geosystems, 20(1), 120–147. https://doi.org/10.1029/2018gc007804
                  Foley, S. F. (1992). Vein‐plus‐wall‐rock melting mechanisms in the lithosphere and the origin of potassic alkaline magmas. Lithos, 28(3–6),
                    435–453. https://doi.org/10.1016/0024‐4937(92)90018‐t
                  Foley, S. F., Ezad, I. S., van der Laan, S. R., & Pertermann, M. (2022). Melting of hydrous pyroxenites with alkali amphiboles in the continental
                    mantle: 1. Melting relations and major element compositions of melts. Geoscience Frontiers, 13(4), 101380. https://doi.org/10.1016/j.gsf.2022.
                    101380
                  Foley, S. F., & Fischer, T. P. (2017). An essential role for continental rifts and lithosphere in the deep carbon cycle. Nature Geoscience, 10(12),
                    897–902. https://doi.org/10.1038/s41561‐017‐0002‐7
                  Foley, S. F., Link, K., Tiberindwa, J., & Barifaijo, E. (2012). Patterns and origin of igneous activity around the Tanzanian craton. Journal of
                    African Earth Sciences, 62(1), 1–18. https://doi.org/10.1016/j.jafrearsci.2011.10.001
                  Foley, S. F., & Pintér, Z. (2018). Primary melt compositions in the Earth’s mantle. In Magmas under pressure (pp. 3–42). Elsevier.
                  Foley, S. F., Yaxley, G., Rosenthal, A., Buhre, S., Kiseeva, E., Rapp, R., & Jacob, D. (2009). The composition of near‐solidus melts of peridotite in
                    the presence of CO2 and H2O between 40 and 60ákbar. Lithos, 112, 274–283. https://doi.org/10.1016/j.lithos.2009.03.020
                  Gardés, E., Gaillard, F., & Tarits, P. (2014). Toward a unified hydrous olivine electrical conductivity law. Geochemistry, Geophysics, Geosystems,
                    15(12), 4984–5000. https://doi.org/10.1002/2014GC005496
                  Giuliani, A., & Pearson, D. G. (2019). Kimberlites: From deep Earth to diamond mines. Elements: An International Magazine of Mineralogy,
                    Geochemistry, and Petrology, 15(6), 377–380. https://doi.org/10.2138/gselements.15.6.377
                  Glover, P. W. (2010). A generalized Archie’s law for n phases. Geophysics, 75(6), E247–E265. https://doi.org/10.1190/1.3509781
                  Grantham, G., Maboko, M., & Eglington, B. (2003). A review of the evolution of the Mozambique Belt and implications for the amalgamation and
                    dispersal of Rodinia and Gondwana. Geological Society, London, Special Publications, 206(1), 401–425. https://doi.org/10.1144/gsl.sp.2003.
                    206.01.19
                  Grégoire, M., Bell, D., & Le Roex, A. (2002). Trace element geochemistry of phlogopite‐rich mafic mantle xenoliths: Their classification and their
                    relationship to phlogopite‐bearing peridotites and kimberlites revisited. Contributions to Mineralogy and Petrology, 142(5), 603–625. https://
                    doi.org/10.1007/s00410‐001‐0315‐8
                  Griffin, W. L., Fisher, N. I., Friedman, J. H., O’Reilly, S. Y., & Ryan, C. G. (2002). Cr‐pyrope garnets in the lithospheric mantle 2. Compositional
                    populations and their distribution in time and space. Geochemistry, Geophysics, Geosystems, 3(12), 1–35. https://doi.org/10.1029/
                    2002gc000298
ÖZAYDIN ET AL.                                                                                                                                             16 of 19
                                                                                                                                                                              15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                                            10.1029/2023GC011191
                  Griffin, W. L., O’Reilly, S. Y., Natapov, L. M., & Ryan, C. G. (2003). The evolution of lithospheric mantle beneath the Kalahari Craton and its
                     margins. Lithos, 71(2–4), 215–241. https://doi.org/10.1016/j.lithos.2003.07.006
                  Griffin, W. L., Ryan, C., O’Reilly, S., Nixon, P., & Win, T. (1991). Trace elements in garnets from Tanzanian kimberlites: Relation to diamond
                     content and tectonic setting. In International kimberlite conference: Extended abstracts (Vol. 5, pp. 145–147).
                  Grütter, H. S., Gurney, J. J., Menzies, A. H., & Winter, F. (2004). An updated classification scheme for mantle‐derived garnet, for use by diamond
                     explorers. Lithos, 77(1–4), 841–857. https://doi.org/10.1016/j.lithos.2004.04.012
                  Hansen, S. E., Nyblade, A. A., & Benoit, M. H. (2012). Mantle structure beneath Africa and Arabia from adaptively parameterized P‐wave
                     tomography: Implications for the origin of Cenozoic Afro‐Arabian tectonism. Earth and Planetary Science Letters, 319, 23–34. https://doi.
                     org/10.1016/j.epsl.2011.12.023
                  Hasterok, D., & Chapman, D. S. (2011). Heat production and geotherms for the continental lithosphere. Earth and Planetary Science Letters,
                     307(1–2), 59–70. https://doi.org/10.1016/j.epsl.2011.04.034
                  Heinson, G., Didana, Y., Soeffky, P., Thiel, S., & Wise, T. (2018). The crustal geophysical signature of a world‐class magmatic mineral system.
                     Scientific Reports, 8(1), 10608. https://doi.org/10.1038/s41598‐018‐29016‐2
                  Hirschmann, M. M., Tenner, T., Aubaud, C., & Withers, A. (2009). Dehydration melting of nominally anhydrous mantle: The primacy of par-
                     titioning. Physics of the Earth and Planetary Interiors, 176(1–2), 54–68. https://doi.org/10.1016/j.pepi.2009.04.001
                  Hirth, G., & Kohlstedf, D. (2003). Rheology of the upper mantle and the mantle wedge: A view from the experimentalists. Geophysical
                     Monograph‐American Geophysical Union, 138, 83–106.
                  Hofmann, C., Courtillot, V., Feraud, G., Rochette, P., Yirgu, G., Ketefo, E., & Pik, R. (1997). Timing of the Ethiopian flood basalt event and
                     implications for plume birth and global change. Nature, 389(6653), 838–841. https://doi.org/10.1038/39853
                  Hübert, J., Whaler, K., & Fisseha, S. (2018). The electrical structure of the central main Ethiopian Rift as imaged by magnetotellurics: Implications
                     for magma storage and pathways. Journal of Geophysical Research: Solid Earth, 123(7), 6019–6032. https://doi.org/10.1029/2017jb015160
                  Hui, H., Peslier, A. H., Rudnick, R. L., Simonetti, A., & Neal, C. R. (2015). Plume‐cratonic lithosphere interaction recorded by water and other
                     trace elements in peridotite xenoliths from the Labait volcano, Tanzania. Geochemistry, Geophysics, Geosystems, 16(6), 1687–1710. https://
                     doi.org/10.1002/2015gc005779
                  Kelbert, A., Meqbel, N., Egbert, G. D., & Tandon, K. (2014). ModEM: A modular system for inversion of electromagnetic geophysical data.
                     Computers & Geosciences, 66, 40–53. https://doi.org/10.1016/j.cageo.2014.01.010
                  Kirkby, A., Czarnota, K., Huston, D. L., Champion, D. C., Doublier, M. P., Bedrosian, P. A., et al. (2022). Lithospheric conductors reveal source
                     regions of convergent margin mineral systems. Scientific Reports, 12(1), 1–10. https://doi.org/10.1038/s41598‐022‐11921‐2
                  Kirkby, A., & Doublier, M. P. (2022). Synthetic magnetotelluric modelling of a regional fault network–implications for survey design and
                     interpretation. Exploration Geophysics, 1–12. https://doi.org/10.1080/08123985.2022.2144212
                  Kirkby, A., Zhang, F., Peacock, J., Hassan, R., & Duan, J. (2019). The MTPy software package for magnetotelluric data analysis and visualisation.
                     Journal of Open Source Software, 4(37), 1358. https://doi.org/10.21105/joss.01358
                  Koornneef, J. M., Davies, G. R., Döpp, S. P., Vukmanovic, Z., Nikogosian, I. K., & Mason, P. R. (2009). Nature and timing of multiple
                     metasomatic events in the sub‐cratonic lithosphere beneath Labait, Tanzania. Lithos, 112, 896–912. https://doi.org/10.1016/j.lithos.2009.
                     04.039
                  Koptev, A., Burov, E., Calais, E., Leroy, S., Gerya, T., Guillou‐Frottier, L., & Cloetingh, S. (2016). Contrasted continental rifting via plume‐craton
                     interaction: Applications to central East African Rift. Geoscience Frontiers, 7(2), 221–236. https://doi.org/10.1016/j.gsf.2015.11.002
                  Lee, C., & Rudnick, R. (1999). Compositionally stratified cratonic lithosphere: Petrology and geochemistry of peridotite xenoliths from the Labait
                     tuff cone, Tanzania. In Proceedings of the 7th international kimberlite conference (pp. 503–521).
                  Le Gall, B., Nonnotte, P., Rolet, J., Benoit, M., Guillou, H., Mousseau‐Nonnotte, M., et al. (2008). Rift propagation at craton margin: Distribution
                     of faulting and volcanism in the North Tanzanian Divergence (East Africa) during Neogene times. Tectonophysics, 448(1–4), 1–19. https://doi.
                     org/10.1016/j.tecto.2007.11.005
                  Lehnert, K., Su, Y., Langmuir, C., Sarbas, B., & Nohl, U. (2000). A global geochemical database structure for rocks. Geochemistry, Geophysics,
                     Geosystems, 1(5). https://doi.org/10.1029/1999gc000026
                  Li, Y., Jiang, H., & Yang, X. (2017). Fluorine follows water: Effect on electrical conductivity of silicate minerals by experimental constraints from
                     phlogopite. Geochimica et Cosmochimica Acta, 217, 16–27. https://doi.org/10.1016/j.gca.2017.08.020
                  Liu, H., Zhu, Q., & Yang, X. (2019). Electrical conductivity of OH‐bearing omphacite and garnet in eclogite: The quantitative dependence on
                     water content. Contributions to Mineralogy and Petrology, 174(7), 1–15. https://doi.org/10.1007/s00410‐019‐1593‐3
                  Maboko, M. A. (2000). Nd and sr isotopic investigation of the Archean–Proterozoic boundary in north eastern Tanzania: Constraints on the nature
                     of Neoproterozoic tectonism in the Mozambique Belt. Precambrian Research, 102(1–2), 87–98. https://doi.org/10.1016/s0301‐9268(00)
                     00060‐7
                  Mana, S., Furman, T., Turrin, B. D., Feigenson, M. D., & Swisher, C. C., III. (2015). Magmatic activity across the East African North Tanzanian
                     divergence zone. Journal of the Geological Society, 172(3), 368–389. https://doi.org/10.1144/jgs2014‐072
                  Manya, S., Kobayashi, K., Maboko, M. A., & Nakamura, E. (2006). Ion microprobe zircon U–Pb dating of the late Archaean metavolcanics and
                     associated granites of the Musoma‐Mara Greenstone Belt, Northeast Tanzania: Implications for the geological evolution of the Tanzania
                     Craton. Journal of African Earth Sciences, 45(3), 355–366. https://doi.org/10.1016/j.jafrearsci.2006.03.004
                  Mattsson, H. B., Nandedkar, R. H., & Ulmer, P. (2013). Petrogenesis of the melilititic and nephelinitic rock suites in the lake natron–engaruka
                     monogenetic volcanic field, northern Tanzania. Lithos, 179, 175–192. https://doi.org/10.1016/j.lithos.2013.07.012
                  Meju, M. A., & Sakkas, V. (2007). Heterogeneous crust and upper mantle across southern Kenya and the relationship to surface deformation as
                     inferred from magnetotelluric imaging. Journal of Geophysical Research, 112(B4), B04103. https://doi.org/10.1029/2005jb004028
                  Muhongo, S., Kröner, A., & Nemchin, A. (2001). Single zircon evaporation and SHRIMP ages for granulite‐facies rocks in the Mozambique Belt
                     of Tanzania. The Journal of Geology, 109(2), 171–189. https://doi.org/10.1086/319240
                  Muirhead, J. D., Fischer, T. P., Oliva, S. J., Laizer, A., van Wijk, J., Currie, C. A., et al. (2020). Displaced cratonic mantle concentrates deep carbon
                     during continental rifting. Nature, 582(7810), 67–72. https://doi.org/10.1038/s41586‐020‐2328‐3
                  Muirhead, J. D., & Kattenhorn, S. A. (2018). Activation of preexisting transverse structures in an evolving magmatic rift in East Africa. Journal of
                     Structural Geology, 106, 1–18. https://doi.org/10.1016/j.jsg.2017.11.004
                  Mulibo, G. D., & Nyblade, A. A. (2013a). Mantle transition zone thinning beneath eastern Africa: Evidence for a whole‐mantle superplume
                     structure. Geophysical Research Letters, 40(14), 3562–3566. https://doi.org/10.1002/grl.50694
                  Mulibo, G. D., & Nyblade, A. A. (2013b). The P and S wave velocity structure of the mantle beneath eastern Africa and the African superplume
                     anomaly. Geochemistry, Geophysics, Geosystems, 14(8), 2696–2715. https://doi.org/10.1002/ggge.20150
                  Naif, S., Selway, K., Murphy, B. S., Egbert, G., & Pommier, A. (2021). Electrical conductivity of the lithosphere‐asthenosphere system. Physics of
                     the Earth and Planetary Interiors, 313(2021), 10661. https://doi.org/10.1016/j.pepi.2021.106661
ÖZAYDIN ET AL.                                                                                                                                                 17 of 19
                                                                                                                                                                             15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                                           10.1029/2023GC011191
                  Novella, D., Frost, D. J., Hauri, E. H., Bureau, H., Raepsaet, C., & Roberge, M. (2014). The distribution of H2O between silicate melt and
                     nominally anhydrous peridotite and the onset of hydrous melting in the deep upper mantle. Earth and Planetary Science Letters, 400, 1–13.
                     https://doi.org/10.1016/j.epsl.2014.05.006
                  Nyblade, A. A. (1997). Heat flow across the East African Plateau. Geophysical Research Letters, 24(16), 2083–2086. https://doi.org/10.1029/
                     97gl01952
                  O’Donnell, J., Adams, A., Nyblade, A., Mulibo, G., & Tugume, F. (2013). The uppermost mantle shear wave velocity structure of eastern Africa
                     from Rayleigh wave tomography: Constraints on rift evolution. Geophysical Journal International, 194(2), 961–978. https://doi.org/10.1093/
                     gji/ggt135
                  O’Donnell, J., Selway, K., Nyblade, A. A., Brazier, R., Tahir, N. E., & Durrheim, R. (2016). Thick lithosphere, deep crustal earthquakes and no
                     melt: A triple challenge to understanding extension in the western branch of the East African Rift. Geophysical Journal International, 204(2),
                     985–998. https://doi.org/10.1093/gji/ggv492
                  O’Reilly, S., & Griffin, W. L. (1996). 4‐D lithosphere mapping: Methodology and examples. Tectonophysics, 262(1–4), 3–18. https://doi.org/10.
                     1016/0040‐1951(96)00010‐8
                  Özaydın, S. (2023). Additional data and models for the article “role of metasomatism in the development of the East African Rift at the northern
                     tanzanian divergence: Insights from 3D magnetotelluric modelling” [Dataset]. Zenodo. In role of metasomatism in the development of the East
                     African Rift at the Northern Tanzanian Divergence: Insights from 3D magnetotelluric modelling Retrieved from https://doi.org/10.5281/
                     zenodo.10393508
                  Özaydın, S., & Selway, K. (2020). MATE: An analysis tool for the interpretation of magnetotelluric models of the mantle. Geochemistry,
                     Geophysics, Geosystems, 21(9), 1–26. https://doi.org/10.1029/2020gc009126
                  Özaydın, S., & Selway, K. (2022). The relationship between kimberlitic magmatism and electrical conductivity anomalies in the mantle.
                     Geophysical Research Letters, 49(18), e2022GL099661. https://doi.org/10.1029/2022gl099661
                  Özaydın, S., Selway, K., & Griffin, W. L. (2021). Are xenoliths from southwestern Kaapvaal Craton representative of the broader mantle?
                     Constraints from magnetotelluric modeling. Geophysical Research Letters, 48(11), e2021GL092570. https://doi.org/10.1029/2021gl092570
                  Özaydın, S., Selway, K., Griffin, W. L., & Moorkamp, M. (2022). Probing the southern African lithosphere with magnetotellurics: 2. Linking
                     electrical conductivity, composition, and tectonomagmatic evolution. Journal of Geophysical Research: Solid Earth, 127(3), e2021JB023105.
                     https://doi.org/10.1029/2021jb023105
                  Padrón‐Navarta, J. A., & Hermann, J. (2017). A subsolidus olivine water solubility equation for the Earth’s upper mantle. Journal of Geophysical
                     Research: Solid Earth, 122(12), 9862–9880. https://doi.org/10.1002/2017JB014510
                  Paterson, M. (1982). The determination of hydroxyl by infrared absorption in quartz, silicate glasses and similar materials. Bulletin de Miner-
                     alogie, 105(1), 20–29. https://doi.org/10.3406/bulmi.1982.7582
                  Peslier, A., Woodland, A., Bell, D., Lazarov, M., & Lapen, T. (2012). Metasomatic control of water contents in the Kaapvaal cratonic mantle.
                     Geochimica et Cosmochimica Acta, 97, 213–246. https://doi.org/10.1016/j.gca.2012.08.028
                  Pintér, Z., Foley, S. F., Yaxley, G. M., Rosenthal, A., Rapp, R. P., Lanati, A. W., & Rushmer, T. (2021). Experimental investigation of the
                     composition of incipient melts in upper mantle peridotites in the presence of CO2 and H2O. Lithos, 396, 106224. https://doi.org/10.1016/j.
                     lithos.2021.106224
                  Plasman, M., Hautot, S., Tarits, P., Gautier, S., Tiberi, C., Le Gall, B., et al. (2019). Lithospheric structure of a transitional magmatic to amagmatic
                     continental rift system—Insights from magnetotelluric and local tomography studies in the North Tanzanian Divergence, East African Rift.
                     Geosciences, 9(11), 462. https://doi.org/10.3390/geosciences9110462
                  Rajaonarison, T. A., Stamps, D. S., & Naliboff, J. (2021). Role of lithospheric buoyancy forces in driving deformation in East Africa from 3D
                     geodynamic modeling. Geophysical Research Letters, 48(6), e2020GL090483. https://doi.org/10.1029/2020gl090483
                  Ramirez, F. D. C., Selway, K., Conrad, C. P., & Lithgow‐Bertelloni, C. (2022). Constraining upper mantle viscosity using temperature and water
                     content inferred from seismic and magnetotelluric data. Journal of Geophysical Research: Solid Earth, 127(8), e2021JB023824. https://doi.org/
                     10.1029/2021jb023824
                  Reiss, M., De Siena, L., & Muirhead, J. D. (2022). The interconnected magmatic plumbing system of the Natron Rift. Geophysical Research
                     Letters, 49(15), e2022GL098922. https://doi.org/10.1029/2022gl098922
                  Ritsema, J., Heijst, H. J. V., & Woodhouse, J. H. (1999). Complex shear wave velocity structure imaged beneath Africa and Iceland. Science,
                     286(5446), 1925–1928. https://doi.org/10.1126/science.286.5446.1925
                  Robertson, K., Heinson, G., & Thiel, S. (2016). Lithospheric reworking at the Proterozoic–Phanerozoic transition of Australia imaged using
                     AusLAMP Magnetotelluric data. Earth and Planetary Science Letters, 452, 27–35. https://doi.org/10.1016/j.epsl.2016.07.036
                  Rooney, T. O. (2020). The Cenozoic magmatism of East Africa: Part III–rifting of the craton. Lithos, 360, 105390. https://doi.org/10.1016/j.lithos.
                     2020.105390
                  Rooney, T. O., Herzberg, C., & Bastow, I. D. (2012). Elevated mantle temperature beneath East Africa. Geology, 40(1), 27–30. https://doi.org/10.
                     1130/g32382.1
                  Ryan, C. G., Griffin, W. L., & Pearson, N. J. (1996). Garnet geotherms: Pressure‐temperature data from Cr‐pyrope garnet xenocrysts in volcanic
                     rocks. Journal of Geophysical Research, 101(B3), 5611–5625. https://doi.org/10.1029/95jb03207
                  Sanislav, I. V., Kolling, S. L., Brayshaw, M., Cook, Y. A., Dirks, P. H., Blenkinsop, T. G., et al. (2015). The geology of the giant Nyankanga gold
                     deposit, Geita Greenstone Belt, Tanzania. Ore Geology Reviews, 69, 1–16. https://doi.org/10.1016/j.oregeorev.2015.02.002
                  Selway, K. (2014). On the causes of electrical conductivity anomalies in tectonically stable lithosphere. Surveys in Geophysics, 35(1), 219–257.
                     https://doi.org/10.1007/s10712‐013‐9235‐1
                  Selway, K. (2015). Negligible effect of hydrogen content on plate strength in East Africa. Nature Geoscience, 8(7), 543–546. https://doi.org/10.
                     1038/ngeo2453
                  Selway, K., & O’Donnell, J. (2019). A small, unextractable melt fraction as the cause for the low velocity zone. Earth and Planetary Science
                     Letters, 517, 117–124. https://doi.org/10.1016/j.epsl.2019.04.012
                  Selway, K., Yi, J., & Karato, S.‐I. (2014). Water content of the Tanzanian lithosphere from magnetotelluric data: Implications for cratonic growth
                     and stability. Earth and Planetary Science Letters, 388, 175–186. https://doi.org/10.1016/j.epsl.2013.11.024
                  Stachel, T., Harris, J. W., & Brey, G. P. (1998). Rare and unusual mineral inclusions in diamonds from Mwadui, Tanzania. Contributions to
                     Mineralogy and Petrology, 132(1), 34–47. https://doi.org/10.1007/s004100050403
                  Stamps, D., Flesch, L., Calais, E., & Ghosh, A. (2014). Current kinematics and dynamics of Africa and the East African Rift System. Journal of
                     Geophysical Research: Solid Earth, 119(6), 5161–5186. https://doi.org/10.1002/2013jb010717
                  Stern, R. J. (1994). Arc assembly and continental collision in the Neoproterozoic East African Orogen: Implications for the consolidation of
                     Gondwanaland. Annual Review of Earth and Planetary Sciences, 22(1), 319–351. https://doi.org/10.1146/annurev.ea.22.050194.001535
ÖZAYDIN ET AL.                                                                                                                                                18 of 19
                                                                                                                                                                        15252027, 2024, 1, Downloaded from https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2023GC011191 by National Health And Medical Research Council, Wiley Online Library on [06/02/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
                 Geochemistry, Geophysics, Geosystems                                                                       10.1029/2023GC011191
                  Ten Grotenhuis, S. M., Drury, M. R., Spiers, C. J., & Peach, C. J. (2005). Melt distribution in olivine rocks based on electrical conductivity
                    measurements. Journal of Geophysical Research, 110(B12), B12201. https://doi.org/10.1029/2004jb003462
                  Thomas, R. J., Spencer, C., Bushi, A. M., Baglow, N., Boniface, N., de Kock, G., et al. (2016). Geochronology of the central Tanzania Craton and
                    its southern and eastern orogenic margins. Precambrian Research, 277, 47–67. https://doi.org/10.1016/j.precamres.2016.02.008
                  Tiberi, C., Gautier, S., Ebinger, C., Roecker, S., Plasman, M., Albaric, J., et al. (2019). Lithospheric modification by extension and magmatism at
                    the craton‐orogenic boundary: North Tanzania Divergence, East Africa. Geophysical Journal International, 216(3), 1693–1710. https://doi.
                    org/10.1093/gji/ggy521
                  Tommasi, A., & Vauchez, A. (2001). Continental rifting parallel to ancient collisional belts: An effect of the mechanical anisotropy of the
                    lithospheric mantle. Earth and Planetary Science Letters, 185(1–2), 199–210. https://doi.org/10.1016/s0012‐821x(00)00350‐2
                  Vauchez, A., Dineur, F., & Rudnick, R. (2005). Microstructure, texture and seismic anisotropy of the lithospheric mantle above a mantle plume:
                    Insights from the Labait volcano xenoliths (Tanzania). Earth and Planetary Science Letters, 232(3–4), 295–314. https://doi.org/10.1016/j.epsl.
                    2005.01.024
                  Wang, Y.‐F., Qin, J.‐Y., Soustelle, V., Zhang, J.‐F., & Xu, H.‐J. (2021). Pyroxene does not always preserve its source hydrogen concentration:
                    Clues from some peridotite xenoliths. Geochimica et Cosmochimica Acta, 292, 382–408. https://doi.org/10.1016/j.gca.2020.10.003
                  Wieser, P., Petrelli, M., Lubbers, J., Wieser, E., Ozaydin, S., Kent, A., & Till, C. (2022). Thermobar: An open‐source Python3 tool for ther-
                    mobarometry and hygrometry. Volcanica, 5(2), 349–384. https://doi.org/10.30909/vol.05.02.349384
                  Withers, A. C., Bureau, H., Raepsaet, C., & Hirschmann, M. M. (2012). Calibration of infrared spectroscopy by elastic recoil detection analysis of
                    H in synthetic olivine. Chemical Geology, 334, 92–98. https://doi.org/10.1016/j.chemgeo.2012.10.002
                  Yang, Y., Ingrin, J., Xia, Q., & Liu, W. (2019). Nature of hydrogen defects in clinopyroxenes from room temperature up to 1000C: Implication for
                    the preservation of hydrogen in the upper mantle and impact on electrical conductivity. American Mineralogist, 104(1), 79–93. https://doi.org/
                    10.2138/am‐2019‐6661
ÖZAYDIN ET AL. 19 of 19