<<

Journal of Geophysical Research: Solid Earth

RESEARCH ARTICLE An Integrated Study of the Eolian Dust in Pelagic 10.1002/2017JB014951 From the North Pacific Based on Environmental Key Points: , Transmission Electron Microscopy, • Ferrimagnetic and antiferromagnetic minerals were identified and Diffuse Reflectance Spectroscopy systematically in North Pacific Ocean pelagic sediments Qiang Zhang1,2 , Qingsong Liu3,4 , Jinhua Li2,4, and Youbin Sun5 • Ferrimagnetic minerals record the overall trend of increased eolian dust 1College of Earth and Planetary Sciences, University of Chinese Academy of Sciences, Beijing, China, 2State Key Laboratory inputs since the late Pliocene 3 • A parameter obtained from diffuse of Lithospheric Evolution, Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing, China, Department 4 reflectance spectroscopy is proposed of Ocean Science and Engineering, Southern University of Science and Technology, Shenzhen, China, Laboratory for to trace eolian dust in pelagic , Qingdao National Laboratory for Marine Science and Technology, Qingdao, China, 5State Key Laboratory sediments of Loess and Quaternary Geology, Institute of Earth Environment, Chinese Academy of Sciences, Xi’an, China

Supporting Information: fi • Supporting Information S1 Abstract Eolian dust is a major terrigenous component in North Paci c Ocean pelagic sediments and is an • Table S1 important recorder of Asian terrestrial environmental evolution and Northern Hemisphere . In order to extract and quantify wind-borne mineral signals and develop a reliable eolian dust Correspondence to: proxy for pelagic sediments, we investigated sediments from Ocean Drilling Program Hole 885A, North Q. Liu, fi [email protected] Paci c Ocean, by integrating results from environmental magnetism, transmission electron microscopy, and diffuse reflectance spectroscopy (DRS). Our results indicate that ferrimagnetic and antiferromagnetic minerals coexist in the sediments. The former includes biogenic , nano-sized single domain (titano) Citation: fi Zhang, Q., Liu, Q., Li, J., & Sun, Y. (2018). magnetite inclusions embedded in silicate hosts, submicron vortex state magnetite, and ultra ne An integrated study of the eolian dust in superparamagnetic maghemite. The latter includes and goethite. Our results indicate that magnetic pelagic sediments from the North properties can be used to determine an overall trend of increased eolian dust inputs, with additional fi Paci c Ocean based on environmental fi fl magnetism, transmission electron signi cant in uence from biogenic and volcanogenic components. In contrast, a newly proposed parameter fl microscopy, and diffuse re ectance RelHm+Gt (combined hematite and goethite concentration) obtained from DRS measurements extracts spectroscopy. Journal of Geophysical precisely the eolian signal and precludes other components. Our DRS results indicate that hematite/goethite – Research: Solid Earth, 123, 3358 3376. fi https://doi.org/10.1002/2017JB014951 and siliceous fractions deposited at Hole 885A were jointly delivered to the North Paci c Ocean from Asian dust sources. Therefore, we suggest that RelHm+Gt is a reliable eolian dust proxy for pelagic sediments in Received 7 SEP 2017 this region. Accepted 23 MAR 2018 Accepted article online 30 MAR 2018 Published online 29 MAY 2018 1. Introduction Eolian dust preserved in terrestrial stratigraphic profiles (An et al., 2001; Guo et al., 2002; Liu, 1985), ice cores (Biscaye et al., 1997; Bory et al., 2003), and marine sediments (Bailey et al., 2011; Larrasoaña et al., 2003; Rea, 1994; Roberts, Rohling, et al., 2011) carries considerable information about the aridification history of source areas and atmospheric circulation over different timescales. The concentration and grain size of eolian dust are linked strongly with environment and . For example, eolian dust fluxes in glacial stages tend to be greater than in interglacial periods (Lambert et al., 2008; Maher et al., 2010). In addition, the grain size of eolian dust can be used to semiquantitatively identify changes in atmospheric circulation intensity (Janecek & Rea, 1985; Sun et al., 2006). Moreover, eolian dust can influence climate and ocean ecosystems directly by changing the properties of the atmosphere (Spracklen et al., 2008) and by influencing the biogeochemistry of (Jickells et al., 2005). Present-day global eolian dust sources are characteristic by a zonal distribution that extends from North Africa to eastern China (Prospero et al., 2002). As shown in satellite images, eolian dust masses can be trans- ported from Asian inland sources to the North Pacific Ocean, with activity greater in springtime (Husar et al., 1997; Prospero et al., 2002). The Taklimakan Desert, Gobi Desert, Hexi Corridor, Chinese Loess Plateau, and Mongolian Plateau are major representative East Asian dust sources (Prospero et al., 2002; Wang et al., 2004). Deep-sea basins with stable sedimentary environments in the North Pacific Ocean lie downwind from the major Asian dust sources (Tanaka & Chiba, 2006) and receive abundant wind-borne minerals that are fl fi ©2018. American Geophysical Union. deposited on the sea oor. Therefore, the North Paci c Ocean is an ideal natural laboratory for eolian dust All Rights Reserved. research. Numerous studies of eolian dust have focused on short-term dust deposition in the North Pacific

ZHANG ET AL. 3358 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Ocean. For instance, quartz abundance is used as an eolian proxy and is coherent with the δ18O record, where more quartz entered the ocean during glacial periods at core RC14-105 (Pisias & Leinen, 1984). Grain size fluc- tuations for eolian components in core KK75-02 were found to be controlled by orbital forcing over the past 700,000 years (Janecek & Rea, 1985). A Quaternary eolian record has also been constructed from core V21-146 on Shatsky Rise, where the maxima and minima of eolian fraction mass accumulation rates correspond exactly to loess and paleosol layers, respectively, in the Xifeng profile, Chinese Loess Plateau (Hovan et al., 1989). The history of eolian over the past 70 million years has also been reconstructed from long cores (e.g., LL44-GPC3 and DSDP Site 576) by Janecek and Rea (1983) and Janecek (1985). A series of deep-sea cores along a transect in the North Pacific Ocean was obtained during Ocean Drilling Program (ODP) Leg 145, which provided the opportunity to investigate the history of Cenozoic terri- genous deposition in the North Pacific Ocean. For example, about 70 m of pelagic post-Cretaceous sediment was cored at ODP Sites 885/886. This site is located downwind of Asia and is sufficiently far from the conti- nent to preclude riverine inputs and ice-rafted debris influences (Snoeckx et al., 1995). Based on isotope analyses, sediments from Site 885/886 have a provenance from the northern Tibetan Plateau and Gobi Desert (Pettke et al., 2000). Snoeckx et al. (1995) recovered a history of eolian deposition over the past 80 million years at Hole 885A. Sun and Liu (2007) reported a new higher-resolution eolian record from this core that spanned the late Pliocene to early Pleistocene and compared it with eolian deposition on the Chinese Loess Plateau. These studies have made significant contributions to understand eolian evolutionary history at different timescales. For these previous studies, the eolian dust signal preserved in pelagic sediments was constructed mainly by

chemical separation procedures, which made use of acetic acid, NaOH (stronger) or Na2CO3 (milder), and sodium dithionite to remove calcium carbonate, opal, and oxides (hydroxides), respectively. However, such chemical procedures are relatively time-consuming and are less suitable for high-resolution studies. Environmental magnetism has also been used to trace eolian signals on the basis of the magnetic properties (type, concentration, and grain size) of wind-borne minerals (Evans & Heller, 2003; Liu et al., 2012; Thompson & Oldfield, 1986). For example, Yamazaki and Ioka (1997) reported that biogenic magnetite produced in situ and detrital magnetite and hematite coexisted in red clay sediments in the North Pacific Ocean. Arnold et al. (1995) analyzed long-term magnetic variations at low resolution on sediments from ODP Sites 885 and 886, and Bailey et al. (2011) also reported magnetic parameters for Hole 885A. Properties of magnetic minerals contained in eolian sediment fractions are primarily controlled by source mineralogy, transportation path- ways, and water regime (Schwertmann & Taylor, 1987). Magnetic parameters of use in dust reconstructions include (deMenocal et al., 1991), hard isothermal remanent magnetization (HIRM; Bloemendal et al., 1988; Doh et al., 1988; Robinson, 1986), and isothermal remanent magnetization (IRM) after

alternating field (AF) demagnetization at 120 mT (IRM900mT@AF120mT; Larrasoaña et al., 2003). However, multiple provenance and post-depositional processes, such as low-temperature oxidation (Johnson & Merrill, 1972; Van Velzen & Dekkers, 1999), reductive dissolution (Roberts, 2015), formation of biogenic magnetite (Hesse, 1994; Roberts, Florindo, et al., 2011), and volcanogenic inputs can cause ambiguities when interpreting the paleoenvironmental significance of magnetic proxies. Diffuse reflectance spectroscopy (DRS) is an alternative way to quantify both hematite and goethite contents (Deaton & Balsam, 1991) and has been used widely in soils (Balsam et al., 2004; Ji et al., 2004; Torrent et al., 2007), aerosols (Shen et al., 2006), and sediments (Balsam & Deaton, 1996; Liu et al., 2011; Oldfield et al., 2014; Zhang et al., 2007). DRS technology has been used, for example, to identify a combination of hematite and goethite in eolian sediment fractions derived from northwest Africa (Balsam et al., 1995). Jiang et al. (2013) and Hu et al. (2016) demonstrated that the combination of DRS and magnetic methods should provide a more reliable approach for determining the concentration of antiferromagnetic minerals in sediments. Previous magnetic studies of Hole 885A have focused on magnetic parameter variations, with less attention paid to the composition and morphology of magnetic minerals at the microscopic scale. Most importantly, there is no demonstrably reliable eolian dust proxy for this hole, which can preclude the influence of other sediment components. Despite demonstration of intensification of Northern Hemisphere glaciations and increased eolian dust flux to the North Pacific Ocean since the late Pliocene revealed by geochemistry and magnetic susceptibility (Haug et al., 2005; Maslin et al., 1998), little attention has been paid to changes in the type, concentration, and grain size (or domain state) of eolian magnetic minerals over this special

ZHANG ET AL. 3359 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

climate transition. Compared with studies of ferrimagnetic minerals (magnetite and maghemite), hematite and goethite have not been well studied. ODP Hole 885A has been well investigated with various methods, which provide a solid paleoclimate background for comparison among different proxies. Based on integrated approaches, we aim to (1) extract wind-borne magnetic minerals from pelagic sediments at ODP Hole 885A; (2) assess changes in type, concentration, and grain size (or domain state) of ferrimagnetic minerals and analyze the corresponding relationship with the history of eolian dust deposition; and (3) identify and quantify hematite/goethite content variations to determine a more reliable eolian dust proxy for the North Pacific Ocean.

2. Materials and Methods In this study, 479 discrete samples were taken from sections 2H-1W to 2H-6W of ODP Hole 885A (44°410N, 168°160W; water depth = 5,708.5 m; Figure 1) between depths of 5.44 and 16.66 mcd (meters composite depth). These samples are composed predominantly of reddish brown and brown clay (with clay concentra- tions of 65–95%; Rea et al., 1993). Sediment age was obtained by linear interpolation on the basis of a magnetic reversal stratigraphy for Hole 885A (Dickens et al., 1995; Weeks et al., 1995). Age control points include the following: Brunhes/Matuyama boundary = 4.86 mcd, top Jaramillo = 5.70 mcd, top Olduvai = 9.57 mcd, onset Olduvai = 10.56 mcd, Matuyama/Gauss = 14.80 mcd, Gauss/Gilbert = 23.70 mcd. Hysteresis loops were measured to maximum applied fields of 1.5 T for samples using a vibrating sample

magnetometer (VSM, Princeton MicroMag 3900). Saturation magnetization (Ms), saturation remanent magne- tization (Mrs or SIRM1.5T), and coercivity (Bc) were obtained after high-field slope correction. The remanence coercivity (Bcr) was obtained by direct current demagnetization of SIRM1.5T back to 1.5 T. The high-field susceptibility (χp) was obtained from the slope of hysteresis loops fitted to data from 0.7 T to 1.4 T. The corrected susceptibility χferri = χlf χp reflects the contributions of ferrimagnetic minerals by removing the influence of paramagnetic minerals (χlf corresponds to the low-frequency susceptibility; Liu et al., 2012). IRM acquisition curves with 100 measurement points and nonlinear field steps on a log scale up to 1.5 T were measured using the VSM. In order to attenuate measurement noise, we used three-point running means to process the measured data. Then IRM acquisition curves were decomposed into magnetic components using the method of Kruiver et al. (2001). First-order reversal curves (FORCs) with 100 FORCs were also measured using the VSM for representative samples. FORC diagrams were processed using the FORCinel software of Harrison and Feinberg (2008). The S ratio and L ratio were obtained from the following procedure: a satura-

tion isothermal remanent magnetization was imparted in a 1.5 T applied field, termed SIRM1.5T, then samples were demagnetized with backfields of 100 and 300 mT, and the corresponding remanence were mea-

sured, termed IRM100mT and IRM300mT. HIRM300T is defined as 0.5 × (SIRM1.5T + IRM300mT), and the S ratio is defined as IRM300mT/SIRM1.5T (King & Channell, 1991). Analogously, HIRM100mT is defined as 0.5 × (SIRM1.5T + IRM100mT). HIRM100mT and HIRM300mT are equivalent to the SIRM after AF demagnetization at 100 (IRM@100mT) and 300 mT (IRM@300mT), respectively (Liu et al., 2007), and the L ratio is defined as HIRM300mT/HIRM100mT (Liu et al., 2007). Low (470 Hz) and high (4700 Hz) frequency magnetic susceptibility and anhysteretic remanent magnetization susceptibility (χARM) data were obtained from Bailey et al. (2011). In order to determine the , low-temperature experiments were conducted using a Quantum Design Magnetic Property Measurement System for representative samples. First, samples were

cooled from 300 to 20 K in zero field, then a SIRM2.5T was imparted at 20 K in a 2.5 T field. The SIRM2.5T was then demagnetized thermally by warming from 20 to 300 K in zero field at a temperature sweep rate of 5 K/min. Temperature dependent susceptibility (χ-T) curves were measured for representative samples using a Kappabridge MFK1-FA magnetic susceptibility meter in an argon environment. Magnetic minerals were extracted from bulk sediment for selected samples and were deposited onto carbon- coated copper grids for transmission electron microscopy (TEM) investigations. TEM experiments were performed with a JEOL 2100 transmission electron microscope operating at 200 kV. Energy-dispersive X- ray spectroscopy (EDXS) analysis was performed to analyze the chemical composition of targets at nanometer scales. DRS was measured for all samples using a Cary 5000 ultraviolet-visible-infrared spectrometer equipped

with BaSO4 as the white standard. The scan rate was 300 nm/min from 300 to 2,600 nm in 0.5 nm

ZHANG ET AL. 3360 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 1. Location of ODP Hole 885A (red circles), Asian inland dust sources (gray white shadow area with number 1 to 6), eolian deposition area in North Pacific Ocean (light-medium-dark yellow area), and trajectories of westerlies (green arrows; Merrill et al., 1989). Asian dust source numbers 1 to 6 indicate Taklimakan, Gobi, Badain Juran, Tengger, Mu Us, Mongolia Gobi desert, or sandy desert, respectively (Sun et al., 2013). “X” with dashed black line indicates southern limit ice-rafted debris (Bigg et al., 2008). The dashed and purple lines indicate the fluxes of mineral aerosol to the (in mg cm 2 ka 1; Duce et al., 1991). The black number in contours within yellow area indicates spatial distribution of <2 μm fraction weight percent of surface samples in the North Pacific Ocean (Arnold, 1996); contour interval is 5%.

steps. DRS data were first transformed into the Kubelka-Munk (K-M) remission function (F(R)=(1 R)2/2R, R is reflectance), and then second derivative curves of F(R) were calculated following Scheinost et al. (1998) and Torrent et al. (2007). Characteristic bands are located at approximately 535 and 425 nm for pure hematite and goethite, respectively, in second derivative curves (Scheinost et al., 1998; Torrent et al.,

2007). The corresponding band intensity for hematite and goethite are defined as IHm and IGt, which are proportional to the concentration of hematite and goethite (Scheinost et al., 1998). We have developed a new proxyRelHm+Gt, which represents the relative concentration of antiferromagnetic I I P Hmi P Gti minerals: RelHmþGt ¼ n n þ n (IHm and IGt indicate the intensity of hematite and goethite, i I I i¼1 Hmi i¼1 Gti and n denotes the number of samples).

3. Results 3.1. Variation of Magnetic Parameters With Depth Environmental magnetic parameters for ODP Hole 885A are shown in Figure 2 from 2.79 to 0.93 Ma. On the basis of trends in these parameters, the section can be divided into three units: Unit 1: 2.58–2.79 Ma, Unit 2: 2.03–2.58 Ma, and Unit 3: 0.93–2.03 Ma. Results are discussed below for each unit, respectively.

Unit 1 (2.58–2.79 Ma): Variations of χferri, χp, χARM, Ms, and SIRM1.5T suggest that the concentration of ferrimag- netic (magnetite/maghemite) and paramagnetic (silicate) minerals decreases first and then gradually

increases upward to the transitional zone (at 2.58 Ma) between Units 1 and 2. Generally, χARM/SIRM variations reflect changes in magnetic grain size (Peters & Dekkers, 2003; Thompson & Oldfield, 1986); however, ARM

may be influenced significantly by interactions among magnetic grains (Sugiura, 1979), so that χARM/ SIRM1.5 may not indicate accurately grain size changes at Hole 885A. Samples from Unit 1 are characterized by higher χARM/SIRM1.5T, which may reflect weaker magnetic interactions where particles acquire a higher ARM. Bcr is sensitive to magnetic grain-size and mineralogy changes. The increase in concentration- dependent magnetic parameters at ~2.58 Ma is associated with a synchronous change in χARM/SIRM1.5T and Bcr values, which indicates that magnetic interactions and high-coercivity magnetic minerals increased at this point. χfd is an indicator of fine magnetic particles near the superparamagnetic (SP)/single domain (SD) threshold (Dearing et al., 1996; Liu et al., 2012). Thus, the lower χfd value in this unit suggests an insignif- icant contribution of SP particles. The HIRM and S ratio can be used to quantify the absolute and relative concentration of antiferromagnetic minerals (hematite and goethite), respectively, only when the L ratio is relatively constant (Liu et al., 2007). These two proxies cannot be used to interpret changes in hematite and goethite contents because of drastic L ratio fluctuations in this unit. In the transition zone between

Units 1 and 2, the L ratio is relatively stable; hence, the change in HIRM300mT value likely indicates

ZHANG ET AL. 3361 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

increased hematite and goethite inputs at 2.58 Ma, which is consistent with previous reports from the low-latitude North Pacific Ocean (Yamazaki & Ioka, 1997).

Unit 2 (2.03–2.58 Ma): Overall, the values of χferri, Ms, SIRM1.5T, and HIRM300mT are relatively higher between 2.03 and 2.58 Ma than in Unit 1. χp is relatively constant, which indicates a stable input of silicates, while χARM and the S ratio decrease, χARM/SIRM1.5T is low, and Bcr increases remarkably compared to Unit 1. It can be inferred that the concentration of particles with stronger magnetic interactions and higher coercivity increases in this unit. Unit 3 (0.93–2.03 Ma): In this unit, four layers (Rea et al., 1993)

have high concentration-dependent parameters and Bcr. Apart from these layers, χferri, χp, Ms, SIRM1.5T, HIRM300mT, and Bcr are all higher than in Unit 2. In general, strongly similar variations among concentration-dependent

magnetic parameters (χferri, Ms, SIRM1.5T, and HIRM300mT) suggest that magnetic enhancement across the transition at 2.58 Ma is dominantly con- trolled by increased ferrimagnetic (magnetite and maghemite) and anti- ferromagnetic mineral (hematite and goethite) contents. 3.2. Hysteresis Data

Plots of Mrs/Ms and Bcr/Bc can provide information about magnetic prop- erty variations in a sample set (Day et al., 1977; Dunlop, 2002a, 2002b). Data from surface samples from Asian inland dust sources (Taklimakan Desert and Qaidam Basin; Liu et al., 2015) extend the trend from the distri-

bution of data points for ODP Hole 885A to lower Mrs/Ms and higher Bcr/Bc values that are indicative of coarser multidomain particles (Figure 3a). The data distribution from finer-grained material at Hole 885A to coarser in Asian inland dust sources (Figure 3a) is interpreted as a response to sorting by wind from source to sink. The overall data trend is from the upper left to lower right, where data for volcanic ash layers diverge from the trend due to abundant SP particles in the ash layers (Figure 3b). 3.3. Bivariate Diagrams for Magnetic Parameters The studied samples are predominantly reddish brown and brown clay (Rea et al., 1993), so the influence on magnetic properties of post- Figure 2. Magnetic parameters versus age of Hole 885A from 2.79 to depositional magnetic mineral oxidation should be considered, especially 0.93 Ma. (a) Ferrimagnetic susceptibility, χferri = χlf χp (χlf, low-frequency transformations from SD magnetite to SP maghemite through oxidation in χ χ susceptibility; p, paramagnetic susceptibility); (b) fd, frequency-dependent pelagic clays (Johnson & Merrill, 1972; Kent & Lowrie, 1974). The absence of susceptibility (χfd = χlf χhf; χlf, low-frequency susceptibility; χhf, correlation between χferri and χfd for non-ash layers may reflect that varia- high-frequency susceptibility), (c) χp, paramagnetic susceptibility; (d) χARM, tions in the amount of SP particles are independent of the total amount of anhysteretic remanent magnetization susceptibility; (e) Ms, saturation magnetization; (f) SIRM1.5T, saturation isothermal remanent magnetization; magnetic minerals (Figure 4a). In contrast to samples without volcanic (g) HIRM300mT, hard remanent magnetization; (h) L ratio; (i) S ratio; (j) Bcr, influences, there is a positive relationship between these two parameters χ remanence coercivity; and (k) ARM/SIRM1.5T. The shallow, medium, and in volcanic ash layers, which indicates that volcanism has generated abun- deep gray shading indicate Unit 1, Unit 2, and Unit 3, respectively. The four fi χ vertical dotted bands depict volcanic ash layers (labeled as Ash 1, Ash 2, dant ultra ne SP particles (Figure 4a). ferri is directly proportional to Ms Ash 3, and Ash 4; Rea et al., 1993). Note that the value of χARM, Ms, SIRM1.5T, (Figure 4b), which demonstrates that the bulk magnetic properties are and HIRM300mT were clipped at volcanic ash layers in order to clarify the controlled mainly by the ferrimagnetic mineral concentration. Indicators variation tendency. The original value of related magnetic parameters is of ferrimagnetic minerals (χferri) and silicates (χp) fall into two groups exhibited in Figure S1 in the supporting information. (Figure 4c). These two parameters are correlated positively especially in Group 1, which suggests that the concentration of ferrimagnetic minerals

and silicates could have a similar origin. HIRM300mT and L ratio have a linear trend for samples in Unit 1 (Figure 4d). Therefore, HIRM300mT variations are affected largely by hematite/goethite coercivity in this unit. Ms for magnetite/maghemite is about two orders of magnitude higher than for hematite/goethite and is

ZHANG ET AL. 3362 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 3. (a) Day plot for samples from Hole 885A, Qaidam Basin, and Taklimakan desert (the latter two are from Liu et al., 2015) and (b) magnified Day plot, dashed green circle separates the samples in volcanic ash layers from other samples.

independent of grain size; hence, Ms provides a measure of the absolute concentration of ferrimagnetic minerals. HIRM300mT is correlated positively with Ms (Figure 4e), which suggests that hematite/goethite and ferrimagnetic mineral have a similar origin.

3.4. Temperature-Dependent Measurements Temperature-dependent magnetic susceptibility (χ-T) measurements can provide information about miner-

alogy and transformation of magnetic minerals by recognizing the Curie temperature (Tc) and other special features (Dunlop & Özdemir, 1997). χ-T curves for selected samples are shown in Figure 5 (Figures 5a-1, 5a-5,

Figure 4. Bivariate diagrams of magnetic parameters. Correlation diagrams between (a) χferri and χfd, (b) χferri and Ms, (c) χferri and χp, (d) HIRM300mT and L ratio, and (e) HIRM300mT and Ms.

ZHANG ET AL. 3363 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 5. Temperature-dependent measurement curves, FORC diagrams, and decomposition of IRM acquisition curves. (a-1, a-5, b-1, b-5, c-1, c-5, d-1) Curves of tem- perature-dependent magnetic susceptibility; (a-2, a-6, b-2, b-6, c-2, c-6, d-2) Low-temperature warming of SIRM curves; (a-3, a-7, b-3, b-7, c-3, c-7, d-3) FORC diagrams, with the smoothing factor of 3; (a-4, a-8, b-4, b-8, c-4, c-8, d-4) decomposition of IRM acquisition curves; the pink, green, deep blue, light blue, and red curves in decomposition diagrams denote Component 1, Component 2, Component 3, Component 4, and sum of components, respectively.

ZHANG ET AL. 3364 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

5b-1, 5b-5, 5c-1, 5c-5, and 5d-1). All heating curves (red) are characterized by a sharp drop at about 585°C,

which is the Tc of magnetite (Evans & Heller, 2003). However, the cooling curves lie above the heating curves due to formation of new magnetite from other minerals (Fe-bearing silicates and clay minerals) during heating (Dunlop & Özdemir, 1997; Hunt et al., 1995). Low-temperature measurements have been widely used to identify mineralogy and grain size (or domain state) of magnetic minerals (Moskowitz et al., 1998; Özdemir et al., 1993). For example, the Verwey (Verwey, 1939) and Morin transitions (Morin, 1950) are key features for identifying magnetite and hematite,

respectively. For stoichiometric magnetite, the Verwey transition temperature (Tv) is near 120 K, and the amplitude of magnetization decrease at Tv is gradually suppressed when domain state changes from the multidomain to the SD state (Dunlop & Özdemir, 1997) because magnetite is usually partially oxidized to

maghemite, so that Tv can be partially or completely suppressed in fine particles (Özdemir et al., 1993). All samples have almost identical curves with no obvious Verwey transition (Figures 5a-2, 5a-6, 5b-2, 5b-6, 5c-2, 5c-6, and 5d-2). This suggests that magnetite particles in the studied samples have undergone significant oxidation, as is the case with other pelagic sediments from the Pacific Ocean (Smirnov & Tarduno, 2000; Yamazaki, 2009; Yamazaki & Ioka, 1997).

3.5. FORC Diagrams and IRM Acquisition Curves FORC diagrams can be used to identify domain state and to discriminate among different components in a magnetic mineral assemblage (Pike et al., 1999; Roberts et al., 2000). In Unit 1, FORC diagrams are character-

ized by elongated contours along the horizontal axis with small vertical spread (Bu < 5 mT; Figures 5a-3 and 5a-7), which indicate typical SD behavior with weak magnetostatic interactions (Roberts et al., 2000). FORC diagrams for samples from Units 2 and 3 (Figures 5b-3, 5b-7, 5c-3, and 5c-7) have diverging outer contours that are indicative of vortex state particles in addition to the non-interacting SD particles indicated by the inner contours (Roberts et al., 2000, 2017). A FORC diagram for a volcanic ash layer contains a secondary peak near the origin of the diagram, with contours nearly parallel to the vertical axis (Figure 5d-3), which is indica- tive of SP particles (Pike et al., 2001; Roberts et al., 2000), along with a main peak due to SD and vortex state

particles. Along the Bc axis, the horizontal contours extend to high fields (i.e., 80–100 mT; Figure 5d-3), which suggests the presence of abundant high-coercivity minerals in volcanic ash layers. IRM acquisition curves can be decomposed into coercivity fractions (Heslop et al., 2002; Kruiver et al., 2001) to provide key information for discriminating components with different magnetic properties. In general, four

components are recognized: Component 1 with B1/2 of 10–18 mT and parameter (DP) of 0.24–0.29; Component 2 with B1/2 of 35–42 mT and DP of 0.20–0.22; Component 3 with B1/2 of 85–141 mT and DP of 0.19–0.28; and Component 4 with B1/2 of 260–380 mT and DP of 0.17–0.21. The details are summar- ized in Table 1. Component 1 may represent ultrafine SP magnetite (maghemite) just above the SP/SD thresh- old size (~20–25 nm), which can have SD behavior because the measurement time is less than the thermal relaxation time (Heslop et al., 2004). Component 2 has a relatively constant mean coercivity and narrow DP and is the major magnetic carrier in Unit 1 (Figures 5a-4 and 5a-8) and is present in all samples. Combined

with the corresponding FORC diagrams (Figures 5a-3 and 5a-7) and higher χARM/SIRM, this component probably represents biogenic SD magnetite, which is present in all samples and has similar magnetic proper- ties as the biogenic magnetite that occurs widely in other Pacific Ocean sediments (Yamazaki, 2009; Yamazaki & Shimono, 2013). Component 3 is characterized by higher coercivity and occurs in significantly higher concentrations in samples from Units 2 and 3 after 2.58 Ma (Figures 5b-4, 5b-8, 5c-4, and 5c-8). This compo- nent is interpreted as eolian maghemite in pelagic North Pacific Ocean sediments and is characterized by high coercivity (~100 mT) and stronger interactions (Yamazaki, 2008, 2009). However, Hu et al. (2013) found a fine pedogenic hematite (~130 mT) in loess/paleosol sequences of the Chinese Loess Plateau. In addition, partially oxidized magnetite can also have elevated coercivity due to the mismatching of unit cells between the maghemite rim and the magnetite core (Cui et al., 1994; Özdemir & O’Reilly, 1982). Component 4 contri- butes <10% to the bulk SIRM and is attributed to antiferromagnetic hematite/goethite.

3.6. TEM Observations TEM observations are used widely to identify magnetic minerals preserved in sediments (Chang et al., 2016; Fassbinder et al., 1990; Maher, 1988). TEM observations from the studied sediments indicate abundant

ZHANG ET AL. 3365 HN TAL. ET ZHANG ora fGohsclRsac:SldEarth Solid Research: Geophysical of Journal

Table 1 The Results of Decomposition of IRM Acquisition Curves for Selected Samples at Hole 885A Component 1 Component 2 Component 3 Component 4

IRM IRM IRM Sample Age IRM B1/2 IRM IRM B1/2 Rel. IRM B1/2 Rel. IRM B1/2 Rel. ID (Ma) Abs. (mT) DP Rel. (%) Abs. (mT) DP (%) Abs. (mT) DP (%) Abs. (mT) DP (%)

885A-2H1W-77-78 1.08 3.90E03 15.49 0.24 14.89 1.10E02 39.81 0.21 41.98 9.80E03 100.00 0.25 37.40 1.50E03 323.59 0.20 5.73 885A-2H1W-101.5-102.5 1.14 3.70E03 17.78 0.28 18.32 9.00E03 40.74 0.22 44.55 6.00E03 100.00 0.22 29.70 1.50E03 269.15 0.21 7.43 885A-2H2W-12-13 1.27 2.60E03 15.85 0.27 18.51 6.45E03 38.90 0.21 45.91 4.45E03 104.71 0.25 31.67 5.50E04 354.81 0.20 3.91 885A-2H2W-71-72 1.40 6.30E03 15.49 0.25 14.35 1.59E02 38.02 0.21 36.22 1.90E02 93.33 0.25 43.28 2.70E03 331.13 0.20 6.15 885A-2H2W-114-115 1.48 4.00E03 17.78 0.28 21.05 8.50E03 40.74 0.22 44.74 5.00E03 100.00 0.22 26.32 1.50E03 269.15 0.21 7.89 885A-2H3W-21-22 1.57 4.90E03 15.85 0.24 14.80 1.35E02 39.81 0.21 40.79 1.25E02 102.33 0.25 37.76 2.20E03 316.23 0.20 6.65 885A-2H3W-65-66 1.66 2.70E03 16.98 0.27 16.29 6.50E03 39.81 0.21 39.23 6.60E03 100.00 0.27 39.83 7.70E04 380.19 0.20 4.65 885A-2H3W-104-105 1.74 6.80E03 17.78 0.28 18.94 1.42E02 41.69 0.22 39.55 1.10E02 100.00 0.22 30.64 3.90E03 269.15 0.21 10.86 885A-2H4W-45-46 1.96 6.00E03 15.14 0.25 13.25 1.90E02 38.90 0.21 41.94 1.70E02 97.72 0.24 37.53 3.30E03 295.12 0.20 7.28 885A-2H4W-62-63 1.99 1.60E02 17.78 0.28 17.80 3.63E02 40.27 0.22 40.38 2.80E02 100.00 0.22 31.15 9.60E03 263.03 0.21 10.68 885A-2H4W-101-102 2.12 2.70E03 17.78 0.28 22.98 5.50E03 38.46 0.22 46.81 2.60E03 100.00 0.21 22.13 9.50E04 275.42 0.20 8.09 885A-2H5W-15-16 2.32 2.30E03 15.85 0.26 21.18 5.50E03 38.02 0.20 50.64 2.80E03 109.65 0.28 25.78 2.60E04 380.19 0.20 2.39 885A-2H5W-42-43 2.40 2.80E03 15.85 0.29 21.37 6.80E03 38.02 0.21 51.91 3.00E03 112.20 0.25 22.90 5.00E04 354.81 0.20 3.82 885A-2H5W-76-77 2.50 5.00E03 17.78 0.28 17.18 1.35E02 41.69 0.22 46.39 7.70E03 100.00 0.21 26.46 2.90E03 260.02 0.20 9.97 885A-2H5W-96-97 2.55 5.90E03 16.60 0.25 18.50 1.30E02 38.90 0.20 40.75 1.16E02 85.11 0.25 36.36 1.40E03 346.74 0.20 4.39 885A-2H5W-103-104 2.57 5.00E03 17.78 0.28 17.41 1.56E02 41.21 0.22 54.39 6.50E03 100.00 0.21 22.63 1.60E03 263.03 0.20 5.57 885A-2H6W-3-4 2.65 5.50E03 12.59 0.29 18.39 1.95E02 38.90 0.22 65.22 3.20E03 114.82 0.20 10.70 1.70E03 295.12 0.20 5.69 885A-2H6W-65-66 2.74 8.30E04 10.00 0.28 12.67 4.70E03 35.48 0.21 71.76 7.00E04 112.20 0.20 10.69 3.20E04 316.23 0.20 4.89 885A-2H6W-77-78 2.76 1.20E03 13.65 0.26 15.79 5.50E03 38.02 0.22 72.37 7.20E04 141.25 0.19 9.47 1.80E04 363.08 0.17 2.37 2 Note. IRM Abs. denotes the absolute value of isothermal remanent magnetization (Am /kg), B1/2 is the field at which half of the SIRM is reached, DP is dispersion parameter, and IRM Rel. denotes the relative percentage content of isothermal remanent magnetization for each component. 10.1002/2017JB014951 3366 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

magnetic minerals with a wide range of grain size and diverse morphologies and can be classified into four types as described below. The first magnetic mineral type is interpreted to be biogenic magnetite (Figures 6a–6d), which has octahedral morphologies and grain size of several tens of nanometers. It is well known that magnetosomes occur abun- dantly in Pacific Ocean pelagic red clays (Yamazaki & Ioka, 1997; Yamazaki & Shimono, 2013). FORC diagrams and IRM acquisition analysis indicate the presence of biogenic magnetite at Hole 885A, which is confirmed by TEM observations. The second magnetic mineral type identified with TEM is nano-sized magnetic mineral inclusions embedded within host silicates, which are present in most selected samples (Figures 6e–6h). The main inclusion micro- structure is characterized by dendritic clusters (Figure 6g) and dendrites (Figures 6e, 6f, and 6h) with mean grain size <100 nm, which is approximately within the size range for SD magnetite. Selected area electron diffraction analysis indicates that these magnetic inclusions have high crystallinity and clear diffraction pat- terns (Figure 6o). EDXS analysis reveals that the inclusions contain abundant Fe and O, with a small quantity of Ti, Cr, Mg, and Co (Figures 6q-2 and 6r-2). The host minerals are rich in O and Si and also have small Al and K contents (Figures 6q-1 and 6r-1). EDXS results for inclusions also indicate Si (e.g., Figure 6q-2) because the inclusions are embedded in host silicate minerals. All of the above morphological, crystallographic, and che- mical compositional characteristics demonstrate that the inclusions are nano-sized (titano) magnetite and that the hosts are likely silicates (e.g., feldspar or pyroxene). The nano-sized magnetic inclusions are adjacent to each other, which may cause significant magnetic interactions that influence ARM values. Magnetic inclu-

sions occur within silicate hosts, so that high-field magnetic susceptibility (χp) indicates magnetic inclusions concentration changes. The third type of magnetic mineral particle identified with TEM occurs mainly in Unit 2 sediments and has submicron scale (>200 nm), subangular morphology (Figures 6k and 6l), and good crystallinity (Figure 6p). This component may be coarser detrital magnetite, which brings vortex state signatures into corresponding FORC diagrams (Figures 5b-3, 5b-7, 5c-3, and 5c-7). In addition to the above three types of magnetic particles, ultra-fine SP maghemite aggregates (Figure 6j) are recognized on the basis of polycrystalline morphological features (Figure 6i). The aggregate maintains the profile of subangular primary coarser vortex state magnetite (Figure 6k), which suggests that SP maghemite might have formed by oxidation of magnetite (Johnson & Merrill, 1972; Kent & Lowrie, 1974).

3.7. Summary of Ferrimagnetic Minerals at Hole 885A Based on the above magnetic and TEM results, the magnetic minerals preserved in the three units of Hole 885A have distinct features as follows: Unit 1 (2.58–2.79 Ma): Based on FORC diagrams, IRM acquisition analysis, and TEM images, biogenic magnetite with minimal magnetostatic interactions is the dominant magnetic carrier in Unit 1 and represents a magnetic background component in all samples. Unit 2 (2.03–2.58 Ma): Coexisting with the biogenic magnetite background, nano-sized (titano) magnetite inclusions

in silicate hosts have significantly increased since 2.58 Ma, as inferred from the increased χp between Units 1 and 2. The relative contribution of Component 3 from IRM analysis also increased since 2.58 Ma (Figures 5b-4, 5b-8, 5c-4, and 5c-8). The fine (titano) magnetite inclusions fall within the SD particle size range and are partially substituted by Ti, Cr, Mg, and Co (Figures 6q-2 and 6r-2). Therefore, this component may have higher coercivity than stoichiometric magnetite. Such inclusions are also adjacent to each other, which may produce

strong magnetic interactions (lower χARM/SIRM). This evidence suggests that Component 3 may represent magnetic inclusions. However, partially oxidized magnetite and fine pedogenic hematite may have coercivity distributions similar to Component 3. Our experimental results may not be enough to identify this compo- nent. Regardless, Component 3 is likely to be of eolian origin. In addition to the above fine particles, coarser detrital vortex state magnetite also increased in abundance at around 2.58 Ma as indicated in FORC diagrams and TEM images. Part of the vortex state magnetite has been transformed into ultrafine SP maghemite, which occurs as polycrystalline aggregates with profiles similar to subangular magnetite, which suggests that it is a product of post-depositional oxidation of magnetite. Unit 3 (0.93–2.03 Ma): Volcanic activity generated abun- dant magnetic minerals and silicates in this unit. Volcanogenic particles are characterized by stronger magnetism and higher coercivity. Regardless of these ash layers, the content of (titano) magnetic inclusions further increased in Unit 3 compared with Unit 2.

ZHANG ET AL. 3367 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 6. Transmission electron microscopy (TEM) images, selected area electron diffraction patterns, and energy-dispersive X-ray spectroscopy (EDXS) analysis. (a-h, j-l) TEM images for magnetic extracts (a, e, f) from samples ODP-885A-2H6W-65-66, (b, d, g, h) from sample ODP-885A-2H6W-23-24, (c, j, k, l) from sample ODP-885A- 2H4W-147-148, (m–p) SAED, (i) polycrystalline diffraction ring for (j), (q, r) EDXS spectra for magnetic inclusions (q-2 and r-2) and their host minerals (q-1 and r-1). The purple cross sign in Figures 6a, 6d, 6h, and 6l indicate the selected area for SAED analysis. Blue and red crosses in panels 6e and 6g indicate the EDXS analysis spots for hosts and inclusions, respectively. The Cu peaks in panels 6q and 6r originate from the TEM grid. Values of Miller indices (hkl) are indicated in panels 6m–6p, which are consistent with those of titanomagnetite or magnetite. The details of the magnetic extraction procedures are exhibited in Text S1 in the supporting information.

ZHANG ET AL. 3368 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 7. Diffuse reflectance spectroscopy (DRS) results. (a, b) Intensity of goethite and hematite obtained from second derivative of K-M remission function from DRS, (c) RelHm+Gt, (d) percentage content of eolian fraction separated by a chemical method from Sun and Liu (2007), (e) HIRM300mT for studied samples at Hole 885A. The shallow, medium, and deep gray shadings indicate Unit 1, Unit 2 and Unit 3, respectively. The four horizontal dotted bands depict the volcanic ash layers (labeled as Ash 1, Ash 2, Ash 3, and Ash 4; Rea et al., 1993). Note that the value of HIRM300mT was clipped at volcanic ash layers in order to clarify the variation tendency. The original value of HIRM300mT is exhibited in Figure S2 in the supporting information.

3.8. DRS Results

Temporal variations of goethite (IGt) and hematite (IHm) intensity obtained from second derivative curves of the DRS K-M remission function, RelHm+Gt, the percentage content of the eolian fraction separated by the chemical method of Sun and Liu (2007), and HIRM300mT are shown in Figure 7. There is a strong similarity in variations of RelHm+Gt (Figure 7c) and the percentage content of eolian dust (Figure 7d), especially in the transition between Units 1 and 2, where they increase synchronously at 2.58 Ma and rise gradually from 1.50 to 0.93 Ma. However, the intensity curves for goethite (Figure 7a) and hematite (Figure 7b) are relatively noisy and do not mirror variations for the eolian component. Eolian sediment fractions often contain a combination of hematite and goethite (Balsam et al., 1995). Liu et al. (2015) inferred that hematite and goethite in surface dust samples in Taklimakan Desert, Qaidam Basin, and Mongolia-Gobi Desert could have

a similar provenance (Figure 8a). IHm and IGt for Hole 885A have a similar distribution pattern (Figure 8b) compared to dust from Asian source regions (Figure 8a), which suggests that hematite and goethite at

Hole 885A are derived from Asian dust sources. HIRM300mT variations (Figure 7e) are consistent with RelHm+Gt (Figure 7c) and percent eolian content (Figure 7d) in Units 1 and 2. However, HIRM300mT values in Unit 3 are influenced significantly by volcanic activity, especially for samples between Ash 2 and Ash 4.

4. Discussion 4.1. Ferrimagnetic Minerals and Eolian Dust Inputs Widespread arid to semiarid regions in northwestern China provide abundant eolian sediment to the North Pacific Ocean by prevailing westerlies (Merrill et al., 1989). Increased ice-rafted debris in deep-sea sediments from the Pacific and Atlantic Oceans reflect intensification of Northern Hemisphere glaciations between 2.74 and 2.54 Ma (Maslin et al., 1998). During this period, the eolian sediment flux to the North Pacific Ocean increased significantly, which indicates intensified aridification in dust source regions and enhanced

ZHANG ET AL. 3369 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Figure 8. Correlation diagrams between intensity of goethite and hematite for samples from (a) Asian inland dust source (data from Liu et al., 2015) and (b) ODP Hole 885A.

atmosphere circulation (Janecek & Rea, 1983; Leinen, 1985). The studied pelagic red clay mainly consists of biogenic, terrigenous, and volcanogenic particles. Increased volcanogenic inputs from the Kamchatka and Aleutian Arcs (Prueher & Rea, 2001) played an important role in enhancing magnetic mineral concentrations in North Pacific Ocean sediments, which produce ambiguities when interpreting eolian signals using magnetic proxies alone. Nevertheless, by combining different magnetic properties, environmental magnetism still provides key information for discriminating sediment components with different origins and for evaluating the degree of volcanogenic particle influences. Hysteresis parameters (Figure 3b) can be an effective tool for discriminating between volcanogenic and terrigenous components and reveal that volcanic ash layers contain abundant high coercivity magnetic minerals with different magnetic behavior from eolian magnetite/maghemite derived from Asia. Environmental magnetism analyses indicate a weaker influence of volcanogenic particles in Units 1 and 2. The magnetic enhancement at 2.58 Ma is mainly controlled by eolian dust inputs from Asia with insignificant volcanogenic inputs. Yamazaki and Ioka (1997) reported that biogenic and detrital magnetite are the major ferrimagnetic compo- nents in North Pacific red clay. The former has lower coercivity and negligible magnetostatic interactions, while the latter has intermediate coercivity and stronger interactions (Yamazaki, 2009). Fine (titano) magne- tite inclusions embedded in silicates have been identified in most of the studied samples. The widespread occurrence of magnetite within silicates in deep-sea sediments has been reported by Chang et al. (2016). Isotopic analysis demonstrates that the silicate fraction at Site 885/886 over the past 11 Ma originated from Asian dust sources (Pettke et al., 2000), which indicates that the fine (titano) magnetite inclusions in silicates

were also derived from Asian inland dust sources. Paramagnetic susceptibility (χp) is often used as an indica- tor of silicates. In this study, χp indicates that magnetite inclusions increased significantly at 2.58 Ma, which suggests further enhanced aridity in Asian dust source regions under intensified Northern Hemisphere glaciations since the late Pliocene (Haug et al., 2005; Janecek & Rea, 1983; Shackleton et al., 1984; Snoeckx et al., 1995; Sun et al., 2006). Environmental magnetism and TEM observations indicate that the change in ferrimagnetic minerals could reflect increased eolian inputs to Hole 885A since the late Pliocene. However, magnetic parameters may

not extract precisely eolian signals because of biogenic and volcanogenic contributions. χp is an ambiguous indicator of eolian magnetite inclusions because volcanic activity also generates abundant silicates. Coarser eolian magnetite could be a more reliable eolian fraction indicator; however, it cannot be identified exclu- sively from magnetic parameters. Therefore, it is difficult to obtain a reliable eolian dust proxy from ferrimag- netic minerals alone. We, therefore, focus below on widespread and stable antiferromagnetic minerals.

4.2. RelHm+Gt From DRS as a Proxy for Eolian Dust Eolian dust typically contains a mixture of magnetically ordered iron oxides and oxyhydroxides, which are sensitive to the process of formation (Schwertmann & Taylor, 1987), transportation (Zhang & Iwasaka, 2004), and deposition (Bailey et al., 2011). However, ferrimagnetic minerals (magnetite/maghemite) are unstable in both oxidizing (Johnson & Merrill, 1972; Kent & Lowrie, 1974) and reducing (Roberts, 2015)

ZHANG ET AL. 3370 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

sedimentary environments. In addition, magnetite/maghemite derived from various sources (e.g., dust, volcanic activity, ice-rafted debris, and magnetosomes) with different magnetic properties may cause ambi- guities in interpreting eolian signals using environmental magnetism. Hematite and goethite are volumetrically dominant magnetic minerals in rocks and sediments (Barron & Montealegre, 1986). Eolian dust from North Africa (Itambi et al., 2009; Larrasoaña et al., 2003, 2008; Lyons et al., 2010; Oldfield et al., 2014; Robinson, 1986), the Asian interior (Arnold et al., 1995; Bailey et al., 2011; Yamazaki, 2009; Yamazaki & Ioka, 1997), and Arabian Peninsula (Bloemendal et al., 1993; Hounslow & Maher, 1999) contains abundant antiferromagnetic minerals (hematite/goethite). Hematite is generally regarded as an end product of mineral weathering process (Barrón et al., 2003). Compared with ferrimagnetic minerals, hematite and goethite are relatively stable in various environments. Hence, antiferro- magnetic minerals are more useful as eolian indicators. HIRM is conventionally employed as a concentration-dependent proxy for hematite and goethite (Thompson & Oldfield, 1986). Doh et al. (1988) found that HIRM variations correlate remarkably well with eolian fluxes in North Pacific pelagic sediments. However, antiferromagnetic minerals may not reach saturation under laboratory magnetic fields, and the accuracy of HIRM may be influenced significantly by high coercivity ferri- magnetic minerals and variations in the coercivity distribution of hematite and goethite (Liu et al., 2007). For example, volcanic ash layers at Hole 885A have high HIRM values, which can be interpreted as due to abun-

dant hematite/goethite. However, the DRS record does not indicate distinct hematite (IHm) and goethite (IGt)

intensity changes within the ash layers. Inconsistencies between HIRM peaks and IHm and IGt minima for volcanic ash layers strongly indicate that, except for hematite and goethite, HIRM can be contaminated by magnetically hard minerals and that DRS parameters can efficiently attenuate the influence of volcanogenic components. Previous studies have demonstrated that high-Ti magnetite is the dominant iron oxide in tephra from Kamchatka (Braitseva et al., 1997; Ponomareva et al., 2015), the Aleutian Arc (Myers et al., 1985; Nye & Turner, 1990; Romick et al., 1990), and in Okhotsk Sea sediments (Wang et al., 2014). In addition, cation sub- stitution in the magnetite crystal lattice may increase coercivity significantly (Akimoto, 1962; Jiang et al., 2016; Liu et al., 2004). Abundant high-coercivity volcanogenic magnetite may disturb the original coercivity distri- bution, so that HIRM may not indicate the concentration of antiferromagnetic minerals in volcanic ash layers. DRS provides a quantitative means to determine the concentration of antiferromagnetic minerals (hematite and goethite) and has a lower detection limit than magnetic measurements, and it is a rapid and nondestruc- tive measurement (Deaton & Balsam, 1991). This method has been used extensively to extract hematite/goethite information in different natural settings (Balsam et al., 2004; Balsam & Deaton, 1996; Ji et al., 2004; Liu et al., 2011; Oldfield et al., 2014; Shen et al., 2006; Torrent et al., 2007; Zhang et al., 2007). The band location and amplitude intensity in DRS spectra are influenced by many factors, such as mathema- tical treatment methods, band overlap, and composition and concentration of minerals. We have adopted the second derivative curves of F(R)=(1 R)2/2R (R is reflectance) to obtain the concentration of hematite and goethite. Compared with first derivative curves, the band location in second derivative curves is not influ- enced by other constituents or by the concentration of minerals in mixtures, and the second derivative is more suitable for natural samples (Liu et al., 2011). Besides natural samples, previous studies have investi- gated synthetic hematite/goethite samples using DRS and traditional rock magnetic methods and found that the DRS band amplitude in second order derivative curves is a relatively reliable concentration proxy for

hematite and goethite (Hu et al., 2016; Jiang et al., 2013; Liu et al., 2011). In this study, we define RelHm+Gt

to combine the concentration of hematite and goethite. The RelHm+Gt record correlates well with percentage content of eolian fractions (Sun & Liu, 2007) from the late Pliocene to early Pleistocene (Figure 7), which further demonstrates that hematite and goethite are weathering products like other silicate fractions and are derived from Asian inland dust sources rather than having volcanogenic or authigenic origins. The DRS proxy succeeds in extracting an eolian signal from ash layers because of insignificant amounts of hematite and goethite in these layers. Compared with chemical separation procedures, DRS measurements are more

rapid and nondestructive. Thus, RelHm+Gt obtained from DRS is a reliable and useful proxy for eolian dust. It is expected that hematite/goethite concentration at Hole 885A may provide detailed information about climate change and the environment of dust sources. Hematite and goethite formation is prevalent under

ZHANG ET AL. 3371 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

opposite conditions, where goethite occurs predominantly in cool and humid , while hematite for- mation is favored in warmer and drier climates (Kämpf & Schwertmann, 1983; Schwertmann, 1993). The major Asian dust sources, Taklimakan Desert (Zhang et al., 1993, 1997), Gobi Desert (Sun et al., 2001), and Qaidam Basin (Pullen et al., 2011), have different altitudes and topographic positions, with temperature and precipitation differences among these regions. Hematite/goethite concentrations and their relative proportions may vary with environmental conditions in source regions. The North Pacific Ocean is an extensive sedimentary sink that receives wind-borne fractions from different Asian dust sources (i.e., multiple-sources-one-sink). Thus, on a hemisphere scale, it is unreasonable to expect antiferromagnetic minerals to reflect details of changes in climate and environments of dust sources. Despite all of this, the combination of hematite and goethite as the major components in eolian fractions is capable of identifying eolian signal in pelagic sediments.

5. Conclusions 1. Four types of ferrimagnetic minerals have been identified in the studied sediments from ODP Hole 885A. The first is biogenic magnetite with low coercivity and negligible magnetostatic interactions, which is the dominant magnetic carrier in Unit 1 (before 2.58 Ma) and exists in all samples. The second is SD (titano) magnetite inclusions embedded in silicate hosts with stronger magnetic interactions and higher coerciv- ity. The third type is submicron and subangular vortex state magnetite that mainly occurs after 2.58 Ma. The fourth type is ultrafine SP maghemite, which occurs as polycrystalline aggregates with subangular magnetite profiles that formed through oxidation of coarse magnetite. 2. Changes in ferrimagnetic minerals could reflect increased eolian dust inputs since the late Pliocene. However, traditional environmental magnetic parameters can be affected severely by biogenic and volca- nogenic magnetic components and limit extraction of reliable eolian signals. 3. Volcanic activity generates abundant strongly magnetic minerals and silicates. Volcanogenic magnetic particles are characterized by stronger magnetizations and higher coercivity but contribute little to antiferromagnetic components. 4. Compared with ferrimagnetic minerals, we find that antiferromagnetic minerals are more suitable for indi- cating eolian dust activity in the studied area. Hematite/goethite and silicate fractions deposited at Hole

885A were delivered to the North Pacific Ocean from Asian dust sources. The parameter RelHm+Gt obtained from DRS measurements correlates remarkably well with the percentage content of eolian fractions and is a more reliable and useful eolian dust proxy than magnetic parameters.

Acknowledgments References The data in this paper can be obtained from the supporting information. This Akimoto, S. (1962). Magnetic properties of FeO-Fe2O3-TiO2 system as a basis of . Journal of the Physical Society of Japan, 17, – study was supported by the National 706 710. Key R&D Program of China An, Z. S., Kutzbach, J. E., Prell, W. L., & Porter, S. C. (2001). Evolution of Asian monsoons and phased uplift of the Himalaya-Tibetan plateau – (2016YFA0601903), the National Natural since late Miocene times. Nature, 411,62 66. Sciences Foundation of China Arnold, E., Leinen, M., & King, J. (1995). Palaeoenvironmental variation based on the mineralogy and rock magnetic properties of sediment fi – (41430962, U1606401), the National from Sites 885 and 886. Proceedings of the Ocean Drilling Program: Scienti c Results, 145, 231 245. fi Program on Global Change and Air-Sea Arnold, E. M. (1996). The rock-magnetic properties, grain size and mineral composition of windborne dust and sediment in the North Paci c Interaction (GASI-GEOGE-03), the Ocean, (Doctoral Thesis). Rhode Island: University of Rhode Island. National Natural Sciences Foundation of Bailey, I., Liu, Q., Swann, G. E., Jiang, Z., Sun, Y., Zhao, X., & Roberts, A. P. (2011). Iron fertilisation and biogeochemical cycles in the sub-Arctic fi fi China (41522402, 4137400), and the Northwest Paci c during the late Pliocene intensi cation of Northern Hemisphere glaciation. Earth and Planetary Science Letters, 307(3-4), – Ocean Drilling Program, which was 253 265. https://doi.org/10.1016/j.epsl.2011.05.029 sponsored by the US National Science Balsam, W., Ji, J., & Chen, J. (2004). Climatic interpretation of the Luochuan and Lingtai loess sections, China, based on changing iron oxide – Foundation (NSF) and participating mineralogy and magnetic susceptibility. Earth and Planetary Science Letters, 223(3-4), 335 348. https://doi.org/10.1016/j.epsl.2004.04.023 countries under management of Joint Balsam, W. L., & Deaton, B. C. (1996). Determining the composition of late Quaternary marine sediments from NUV, VIS, and NIR diffuse fl – – Oceanographic Institutions (JOI), Inc. re ectance spectra. Marine Geology, 134(1 2), 31 55. https://doi.org/10.1016/0025-3227(96)00037-0 The authors thank Heng Zhang for his Balsam, W. L., Otto-Bliesner, B. L., & Deaton, B. C. (1995). Modern and last glacial maximum eolian sedimentation patterns in the Atlantic – help with extracting magnetic minerals Ocean interpreted from sediment iron oxide content. Paleoceanography, 10(3), 493 507. https://doi.org/10.1029/95PA00421 and the TEM engineers at the Institute Barron, V., & Montealegre, L. (1986). Iron oxides and color of Triassic sediments; Application of the Kubelka-Munk theory. American Journal of – of Geology and Geophysics, Chinese Science, 286(10), 792 802. https://doi.org/10.2475/ajs.286.10.792 Academy of Sciences (Lixin Gu and Xu Barrón, V., Torrent, J., & De Grave, E. (2003). Hydromaghemite, an intermediate in the hydrothermal transformation of 2-line ferrihydrite into – – Tang), for their efforts to maintain hematite. American Mineralogist, 88(11 12), 1679 1688. https://doi.org/10.2138/am-2003-11-1207 fi smooth operation of their TEM facility. Bigg, G. R., Clark, C. D., & Hughes, A. L. C. (2008). A last glacial ice sheet on the Paci c Russian coast and catastrophic change arising from coupled ice–volcanic interaction. Earth and Planetary Science Letters, 265(3-4), 559–570. https://doi.org/10.1016/j.epsl.2007.10.052 Biscaye, P. E., Grousset, F. E., Revel, M., Van der Gaast, S., Zielinski, G. A., Vaars, A., & Kukla, G. (1997). Asian provenance of glacial dust (stage 2) in the Greenland Ice Sheet Project 2 ice core, Summit, Greenland. Journal of Geophysical Research, 102(C12), 26,765–26,781. https://doi. org/10.1029/97JC01249

ZHANG ET AL. 3372 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Bloemendal, J., King, J. W., Hunt, A., Demenocal, P. B., & Hayashida, A. (1993). Origin of the sedimentary magnetic record at Ocean Drilling Program sites on the Owen Ridge, western Arabian Sea. Journal of Geophysical Research, 98(B3), 4199–4219. https://doi.org/10.1029/ 92JB02914 Bloemendal, J., Lamb, B., & King, J. (1988). Palaeoenvironmental implications of rock-magnetic properties of late Quaternary sediment cores from the eastern equatorial Atlantic. Paleoceanography, 3(1), 61–87. https://doi.org/10.1029/PA003i001p00061 Bory, A. J. M., Biscaye, P. E., & Grousset, F. E. (2003). Two distinct seasonal Asian source regions for mineral dust deposited in Greenland (NorthGRIP). Geophysical Research Letters, 30(4), 1167. https://doi.org/10.1029/2002GL016446 Braitseva, O. A., Ponomareva, V. V., Sulerzhitsky, L. D., Melekestsev, I. V., & Bailey, J. (1997). Holocene key-marker tephra layers in Kamchatka, Russia. Quaternary Research, 47(02), 125–139. https://doi.org/10.1006/qres.1996.1876 Chang, L., Roberts, A. P., Heslop, D., Hayashida, A., Li, J., Zhao, X., et al. (2016). Widespread occurrence of silicate-hosted magnetic mineral inclusions in marine sediments and their contribution to paleomagnetic recording. Journal of Geophysical Research: Solid Earth, 121, 8415–8431. https://doi.org/10.1002/2016JB013109 Cui, Y., Verosub, K. L., & Roberts, A. P. (1994). The effect of low-temperature oxidation on large multi-domain magnetite. Geophysical Research Letters, 21(9), 757–760. https://doi.org/10.1029/94GL00639 Day, R., Fuller, M., & Schmidt, V. A. (1977). Hysteresis properties of titanomagnetites: Grain-size and compositional dependence. Physics of the Earth and Planetary Interiors, 13(4), 260–267. https://doi.org/10.1016/0031-9201(77)90108-X Dearing, J. A., Hay, K. L., Baban, S. M. J., Huddleston, A. S., Wellington, E. M. H., & Loveland, P. (1996). Magnetic susceptibility of soil: An evaluation of conflicting theories using a national data set. Geophysical Journal International, 127(3), 728–734. https://doi.org/10.1111/ j.1365-246X.1996.tb04051.x Deaton, B. C., & Balsam, W. L. (1991). Visible spectroscopy; A rapid method for determining hematite and goethite concentration in geological materials. Journal of Sedimentary Research, 61(4), 628–632. https://doi.org/10.1306/D4267794-2B26-11D7- 8648000102C1865D deMenocal, P., Bloemendal, J., & King, J. W. (1991). A rock-magnetic record of monsoonal dust deposition to the Arabian Sea: Evidence for a shift in the mode of deposition at 2.4 Ma. Proceedings of the Ocean Drilling Program: Scientific Results, 117, 389–407. Dickens, G. R., Snoeckx, H., Arnold, E., Morley, J. J., Owen, R. M., Rea, D. K., & Ingram, L. (1995). Composite depth scale and stratigraphy for Sites 885/886. Proceedings of the Ocean Drilling Program: Scientific Results, 145, 205–217. Doh, S. J., King, J. W., & Leinen, M. (1988). A rock-magnetic study of giant piston core LL44-GPC3 from the central North Pacific and its paleoceanographic implications. Paleoceanography, 3(1), 89–111. https://doi.org/10.1029/PA003i001p00089 Duce, R. A., Liss, P. S., Merrill, J. T., Atlas, E. L., Buat-Menard, P., Hicks, B. B., et al. (1991). The atmospheric input of trace species to the world ocean. Global Biogeochemical Cycles, 5(3), 193–259. https://doi.org/10.1029/91GB01778 Dunlop, D. J. (2002a). Theory and application of the Day plot (Mrs/Ms versus Hcr/Hc) 1. Theoretical curves and tests using titanomagnetite data. Journal of Geophysical Research, 107(B3), 2056. https://doi.org/10.1029/2001JB000486 Dunlop, D. J. (2002b). Theory and application of the Day plot (Mrs/Ms versus Hcr/Hc) 2. Application to data for rocks, sediments, and soils. Journal of Geophysical Research, 107(B3), 2057. https://doi.org/10.1029/2001JB000487 Dunlop, D. J., & Özdemir, Ö. (1997). Rock magnetism: Fundamentals and frontiers. New York: Cambridge University Press. https://doi.org/ 10.1017/CBO9780511612794 Evans, M. E., & Heller, F. (2003). Environmental magnetism: Principles and applications of enviromagnetics. San Diego, CA: Academic Press. Fassbinder, J. W., Stanjekt, H., & Vali, H. (1990). Occurrence of magnetic bacteria in soil. Nature, 343(6254), 161–163. https://doi.org/10.1038/ 343161a0 Guo, Z. T., Ruddiman, W. F., Hao, Q. Z., Wu, H. B., Qiao, Y. S., Zhu, R. X., et al. (2002). Onset of Asian desertification by 22 Myr ago inferred from loess deposits in China. Nature, 416(6877), 159–163. https://doi.org/10.1038/416159a Harrison, R. J., & Feinberg, J. M. (2008). FORCinel: An improved algorithm for calculating first-order reversal curve distributions using locally weighted regression smoothing. Geochemistry, Geophysics, Geosystems, 9, Q05016. https://doi.org/10.1029/2008GC001987 Haug, G. H., Ganopolski, A., Sigman, D. M., & Rosell-Mele, A. (2005). North Pacific seasonality and the glaciation of North America 2.7 million years ago. Nature, 433(7028), 821–825. https://doi.org/10.1038/nature03332 Heslop, D., Dekkers, M. J., Kruiver, P. P., & Van Oorschot, I. H. M. (2002). Analysis of isothermal remanent magnetization acquisition curves using the expectation-maximization algorithm. Geophysical Journal International, 148(1), 58–64. https://doi.org/10.1046/j.0956- 540x.2001.01558.x Heslop, D., McIntosh, G., & Dekkers, M. J. (2004). Using time-and temperature-dependent Preisach models to investigate the limitations of modelling isothermal remanent magnetization acquisition curves with cumulative log Gaussian functions. Geophysical Journal International, 157(1), 55–63. https://doi.org/10.1111/j.1365-246X.2004.02155.x Hesse, P. P. (1994). Evidence for bacterial palaeoecological origin of mineral magnetic cycles in oxic and sub-oxic Tasman Sea sediments. Marine Geology, 117(1–4), 1–17. https://doi.org/10.1016/0025-3227(94)90003-5 Hounslow, M. W., & Maher, B. A. (1999). Source of the climate signal recorded by magnetic susceptibility variations in Indian Ocean sedi- ments. Journal of Geophysical Research, 104(B3), 5047–5061. https://doi.org/10.1029/1998JB900085 Hovan, S. A., Rea, D. K., Pisias, N. G., & Shackleton, N. J. (1989). A direct link between the China loess and marine δ18O records: Aeolian flux to the North Pacific. Nature, 340(6231), 296–298. https://doi.org/10.1038/340296a0 Hu, P., Jiang, Z., Liu, Q., Heslop, D., Roberts, A. P., Torrent, J., & Barrón, V. (2016). Estimating the concentration of aluminum-substituted hematite and goethite using diffuse reflectance spectrometry and rock magnetism: Feasibility and limitations. Journal of Geophysical Research: Solid Earth, 121, 4180–4194. https://doi.org/10.1002/2015JB012635 Hu, P., Liu, Q., Torrent, J., Barrón, V., & Jin, C. (2013). Characterizing and quantifying iron oxides in Chinese loess/paleosols: Implications for pedogenesis. Earth and Planetary Science Letters, 369–370(3), 271–283. Hunt, C. P., Moskowitz, B. M., & Banerjee, S. K. (1995). Magnetic properties of rocks and minerals. In Rock physics & phase relations: A handbook of physical constants (Vol. 3, pp. 189–204). Washington, DC: American Geophysical Union. https://doi.org/10.1029/RF003p0189 Husar, R. B., Prospero, J. M., & Stowe, L. L. (1997). Characterization of tropospheric aerosols over the oceans with the NOAA advanced very high resolution radiometer optical thickness operational product. Journal of Geophysical Research, 102(D14), 16,889–16,909. https://doi. org/10.1029/96JD04009 Itambi, A. C., Von Dobeneck, T., Mulitza, S., Bickert, T., & Heslop, D. (2009). Millennial-scale northwest African droughts related to Heinrich events and Dansgaard-Oeschger cycles: Evidence in marine sediments from offshore Senegal. Paleoceanography, 24, PA1205. https://doi. org/10.1029/2007PA001570 Janecek, T. R. (1985). Eolian sedimentation in the northwest Pacific Ocean: Preliminary examination of the data from Drilling Project Sites 576 and 578. Initial Report Deep Sea Drilling Project, 86, 589–603. College Station, TX.

ZHANG ET AL. 3373 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Janecek, T. R., & Rea, D. K. (1983). Eolian deposition in the northeast Pacific Ocean: Cenozoic history of atmospheric circulation. Geological Society of America Bulletin, 94(6), 730–738. https://doi.org/10.1130/0016-7606(1983)94%3C730:EDITNP%3E2.0.CO;2 Janecek, T. R., & Rea, D. K. (1985). Quaternary fluctuations in the Northern Hemisphere trade winds and westerlies. Quaternary Research, 24(02), 150–163. https://doi.org/10.1016/0033-5894(85)90002-X Ji, J., Chen, J., Balsam, W., Lu, H., Sun, Y., & Xu, H. (2004). High resolution hematite/goethite records from Chinese loess sequences for the last glacial-interglacial cycle: Rapid climatic response of the East Asian Monsoon to the tropical Pacific. Geophysical Research Letters, 31, L03207. https://doi.org/10.1029/2003GL018975 Jiang, Z., Liu, Q., Colombo, C., Barrón, V., Torrent, J., & Hu, P. (2013). Quantification of Al-goethite from diffuse reflectance spectroscopy and magnetic methods. Geophysical Journal International, 196(1), 131–144. Jiang, Z., Liu, Q., Zhao, X., Roberts, A. P., Heslop, D., Barrón, V., & Torrent, J. (2016). Magnetism of Al-substituted magnetite reduced from Al- hematite. Journal of Geophysical Research: Solid Earth, 121, 4195–4210. https://doi.org/10.1002/2016JB012863 Jickells, T. D., An, Z. S., Andersen, K. K., Baker, A. R., Bergametti, G., Brooks, N., et al. (2005). Global iron connections between desert dust, ocean biogeochemistry, and climate. Science, 308(5718), 67–71. https://doi.org/10.1126/science.1105959 Johnson, H. P., & Merrill, R. T. (1972). Magnetic and mineralogical changes associated with low-temperature oxidation of magnetite. Journal of Geophysical Research, 77(2), 334–341. https://doi.org/10.1029/JB077i002p00334 Kämpf, N., & Schwertmann, U. (1983). Goethite and hematite in a climosequence in southern Brazil and their application in classification of kaolinitic soils. Geoderma, 29(1), 27–39. https://doi.org/10.1016/0016-7061(83)90028-9 Kent, D. V., & Lowrie, W. (1974). Origin of magnetic instability in sediment cores from the central North Pacific. Journal of Geophysical Research, 79(20), 2987–3000. https://doi.org/10.1029/JB079i020p02987 King, J. W., & Channell, J. E. T. (1991). Sedimentary magnetism, environmental magnetism and magnetostratigraphy. Reviews of Geophysics, 29(S1), 358–370. https://doi.org/10.1002/rog.1991.29.s1.358 Kruiver, P. P., Dekkers, M. J., & Heslop, D. (2001). Quantification of magnetic coercivity components by the analysis of acquisition curves of isothermal remanent magnetisation. Earth and Planetary Science Letters, 189(3-4), 269–276. https://doi.org/10.1016/S0012-821X(01)00367-3 Lambert, F., Delmonte, B., Petit, J. R., Bigler, M., Kaufmann, P. R., Hutterli, M. A., et al. (2008). Dust-climate couplings over the past 800,000 years from the EPICA dome C ice core. Nature, 452(7187), 616–619. https://doi.org/10.1038/nature06763 Larrasoaña, J. C., Roberts, A. P., & Rohling, E. J. (2008). Magnetic susceptibility of eastern Mediterranean marine sediments as a proxy for Saharan dust supply? Marine Geology, 254(3-4), 224–229. https://doi.org/10.1016/j.margeo.2008.06.003 Larrasoaña, J. C., Roberts, A. P., Rohling, E. J., Winklhofer, M., & Wehausen, R. (2003). Three million years of monsoon variability over the northern Sahara. Climate Dynamics, 21(7–8), 689–698. https://doi.org/10.1007/s00382-003-0355-z Leinen, M. (1985). Quartz content of northwest Pacific Hole-576a and implications for Cenozoic eolian transport. Initial Reports of the Deep Sea Drilling Project, 86(NOV), 581–588. Liu, Q., Banerjee, S. K., Jackson, M. J., Deng, C., Pan, Y., & Rixiang, Z. (2004). New insights into partial oxidation model of and thermal alteration of magnetic mineralogy of the Chinese loess in air. Geophysical Journal International, 158(2), 506–514. https://doi.org/ 10.1111/j.1365-246X.2004.02348.x Liu, Q., Roberts, A. P., Larrasoana, J. C., Banerjee, S. K., Guyodo, Y., Tauxe, L., & Oldfield, F. (2012). Environmental magnetism: Principles and applications. Reviews of Geophysics, 50, RG4002. https://doi.org/10.1029/2012RG000393 Liu, Q., Roberts, A. P., Torrent, J., Horng, C. S., & Larrasoana, J. C. (2007). What do the HIRM and S-ratio really measure in environmental magnetism? Geochemistry, Geophysics, Geosystems, 8, Q09011. https://doi.org/10.1029/2007GC001717 Liu, Q., Sun, Y., Qiang, X., Tada, R., Hu, P., Duan, Z., et al. (2015). Characterizing magnetic mineral assemblages of surface sediments from major Asian dust sources and implications for the Chinese loess magnetism. Earth, Planets and Space, 67(1), 1–17. Liu, Q. S., Torrent, J., Barrón, V., Duan, Z. Q., & Bloemendal, J. (2011). Quantification of hematite from the visible diffuse reflectance spectrum: Effects of aluminium substitution and grain morphology. Clay Minerals, 46(1), 137–147. https://doi.org/10.1180/claymin.2011.046.1.137 Liu, T. S. (1985). Loess and the Environment. Beijing: China Ocean Press. Lyons, R., Oldfield, F., & Williams, E. (2010). Mineral magnetic properties of surface soils and sands across four North African transects and links to climatic gradients. Geochemistry, Geophysics, Geosystems, 11, Q08023. https://doi.org/10.1029/2010GC003183 Maher, B. A. (1988). Magnetic properties of some synthetic sub-micron magnetites. Geophysical Journal International, 94(1), 83–96. https:// doi.org/10.1111/j.1365-246X.1988.tb03429.x Maher, B. A., Prospero, J. M., Mackie, D., Gaiero, D., Hesse, P. P., & Balkanski, Y. (2010). Global connections between aeolian dust, climate and ocean biogeochemistry at the present day and at the last glacial maximum. Earth-Science Reviews, 99(1-2), 61–97. https://doi.org/10.1016/ j.earscirev.2009.12.001 Maslin, M. A., Li, X. S., Loutre, M. F., & Berger, A. (1998). The contribution of orbital forcing to the progressive intensification of Northern Hemisphere glaciation. Quaternary Science Reviews, 17(4–5), 411–426. https://doi.org/10.1016/S0277-3791(97)00047-4 Merrill, J. T., Uematsu, M., & Bleck, R. (1989). Meteorological analysis of long range transport of mineral aerosols over the North Pacific. Journal of Geophysical Research, 94(D6), 8584–8598. https://doi.org/10.1029/JD094iD06p08584 Morin, F. J. (1950). Magnetic susceptibility of αFe2O3, and αFe2O3 with added titanium. Physical Review, 78(6), 819–820. https://doi.org/ 10.1103/PhysRev.78.819.2 Moskowitz, B. M., Jackson, M., & Kissel, C. (1998). Low-temperature magnetic behavior of titanomagnetites. Earth and Planetary Science Letters, 157(3–4), 141–149. https://doi.org/10.1016/S0012-821X(98)00033-8 Myers, J. D., Marsh, B. D., & Sinha, A. K. (1985). Strontium isotopic and selected trace element variations between two aleutian volcanic centers (Adak and Atka): Implications for the development of arc volcanic plumbing systems. Contributions to Mineralogy and Petrology, 91(3), 221–234. https://doi.org/10.1007/BF00413349 Nye, C. J., & Turner, D. L. (1990). Petrology, geochemistry, and age of the spurr volcanic complex, eastern aleutian arc. Bulletin of Volcanology, 52(3), 205–226. https://doi.org/10.1007/BF00334805 Oldfield, F., Chiverrell, R. C., Lyons, R., Williams, E., Shen, Z., Bristow, C., et al. (2014). Discriminating dusts and dusts sources using magnetic properties and hematite: Goethite ratios of surface materials and dust from North Africa, the Atlantic and Barbados. Aeolian Research, 13, 91–104. https://doi.org/10.1016/j.aeolia.2014.03.010 Özdemir, Ö., Dunlop, D. J., & Moskowitz, B. M. (1993). The effect of oxidation on the Verwey transition in magnetite. Geophysical Research Letters, 20(16), 1671–1674. https://doi.org/10.1029/93GL01483 Özdemir, Ö., & O’Reilly, W. (1982). Magnetic hysteresis properties of synthetic monodomain titanomaghemites. Earth and Planetary Science Letters, 57(2), 437–447. https://doi.org/10.1016/0012-821X(82)90162-5 Peters, C., & Dekkers, M. J. (2003). Selected room temperature magnetic parameters as a function of mineralogy, concentration and grain size. Physics and Chemistry of the Earth, Parts A/B/C, 28(16-19), 659–667. https://doi.org/10.1016/S1474-7065(03)00120-7

ZHANG ET AL. 3374 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Pettke, T., Halliday, A. N., Hall, C. M., & Rea, D. K. (2000). Dust production and deposition in Asia and the North Pacific Ocean over the past 12 Myr. Earth and Planetary Science Letters, 178(3-4), 397–413. https://doi.org/10.1016/S0012-821X(00)00083-2 Pike, C. R., Roberts, A. P., & Verosub, K. L. (1999). Characterizing interactions in fine magnetic particle systems using first order reversal curves. Journal of Applied Physics, 85(9), 6660–6667. https://doi.org/10.1063/1.370176 Pike, C. R., Roberts, A. P., & Verosub, K. L. (2001). FORC diagrams and thermal relaxation effects in magnetic particles. Geophysical Journal International, 145(3), 721–730. https://doi.org/10.1046/j.0956-540x.2001.01419.x Pisias, N. G., & Leinen, M. (1984). Milankovitch forcing of the oceanic system: Evidence from the northwest Pacific. In A. Berger, et al. (Eds.), Milankovitch and climate, NATO ASI Series C (Vol. 126, pp. 307–330). Dordrecht: Reidel. Ponomareva, V., Portnyagin, M., Pevzner, M., Blaauw, M., Kyle, P., & Derkachev, A. (2015). Tephra from andesitic Shiveluch volcano, Kamchatka, NW Pacific: Chronology of explosive eruptions and geochemical fingerprinting of volcanic glass. International Journal of Earth Sciences, 104(5), 1459–1482. https://doi.org/10.1007/s00531-015-1156-4 Prospero, J. M., Ginoux, P., Torres, O., Nicholson, S. E., & Gill, T. E. (2002). Environmental characterization of global, sources of atmospheric soil dust, identified with the Nimbus 7 Total Ozone Mapping Spectrometer (TOMS) absorbing aerosol product. Reviews of Geophysics, 40(1), 1002. https://doi.org/10.1029/2000RG000095 Prueher, L. M., & Rea, D. K. (2001). Tephrochronology of the Kamchatka–Kurile and Aleutian arcs: Evidence for volcanic episodicity. Journal of Volcanology and Geothermal Research, 106(1-2), 67–84. https://doi.org/10.1016/S0377-0273(00)00266-3 Pullen, A., Kapp, P., McCallister, A. T., Chang, H., Gehrels, G. E., Garzione, C. N., et al. (2011). Qaidam Basin and northern Tibetan Plateau as dust sources for the Chinese Loess Plateau and paleoclimatic implications. Geology, 39(11), 1031–1034. https://doi.org/10.1130/G32296.1 Rea, D. K. (1994). The paleoclimatic record provided by eolian deposition in the deep sea: The geologic history of wind. Reviews of Geophysics, 32(2), 159–195. https://doi.org/10.1029/93RG03257 Rea, D. K., Basov, I. A., Janecek, T. R., Palmer-Julson, A., et al. (1993). Sites 885/886. In Proceedings of the Ocean Drilling Program: Initial reports (Vol. 145, pp. 303–334). College Station, TX: Ocean Drilling Program. https://doi.org/10.2973/odp.proc.ir.145.1993 Roberts, A. P. (2015). Magnetic mineral diagenesis. Earth-Science Reviews, 151,1–47. https://doi.org/10.1016/j.earscirev.2015.09.010 Roberts, A. P., Almeida, T. P., Church, N. S., Harrison, R. J., Heslop, D., Li, Y., et al. (2017). Resolving the origin of pseudo-single domain magnetic behavior. Journal of Geophysical Research: Solid Earth, 122, 9534–9558. https://doi.org/10.1002/2017JB014860 Roberts, A. P., Florindo, F., Villa, G., Liao, C., Jovane, L., Bohaty, S. M., et al. (2011). Magnetotactic bacterial abundance in pelagic marine environments is limited by organic carbon flux and availability of dissolved iron. Earth and Planetary Science Letters, 310(3–4), 441–452. https://doi.org/10.1016/j.epsl.2011.08.011 Roberts, A. P., Pike, C. R., & Verosub, K. L. (2000). First-order reversal curve diagrams: A new tool for characterizing the magnetic properties of natural samples. Journal of Geophysical Research, 105(B12), 28,461–28,475. https://doi.org/10.1029/2000JB900326 Roberts, A. P., Rohling, E. J., Grant, K. M., Larrasoaña, J. C., & Liu, Q. (2011). Atmospheric dust variability from Arabia and China over the last 500,000 years. Quaternary Science Reviews, 30(25-26), 3537–3541. https://doi.org/10.1016/j.quascirev.2011.09.007 Robinson, S. G. (1986). The late Pleistocene palaeoclimatic record of North Atlantic deep-sea sediments revealed by mineral-magnetic measurements. Physics of the Earth and Planetary Interiors, 42(1–2), 22–47. https://doi.org/10.1016/S0031-9201(86)80006-1 Romick, J. D., Perfit, M. R., Swanson, S. E., & Shuster, R. D. (1990). Magmatism in the eastern Aleutian arc: Temporal characteristic of igneous activity on Akutan Island. Contributions to Mineralogy and Petrology, 104(6), 700–721. https://doi.org/10.1007/BF01167288 Scheinost, A. C., Chavernas, A., Barrón, V., & Torrent, J. (1998). Use and limitations of second-derivative diffuse reflectance spectroscopy in the visible to near-infrared range to identify and quantity Fe oxide minerals in soils. Clays and Clay Minerals, 46(5), 528–536. https://doi.org/ 10.1346/CCMN.1998.0460506 Schwertmann, U. (1993). Relations between iron oxides, soil color, and soil formation. In J. M. Bigham & E. J. Ciolkosz (Eds.), Soil color (pp. 51–70). Madison: Soil Science Society of America. Schwertmann, U., & Taylor, R. M. (1987). Iron oxides. In J. B. Dixon & S. B. Weed (Eds.), Minerals in soil environments (2nd ed., pp. 379–438). Madison, WI: Soil Science Society of America. Shackleton, N. J., Backman, J., Zimmerman, H. T., Kent, D. V., Hall, M. A., Roberts, D. G., et al. (1984). Oxygen isotope calibration of the onset of ice-rafting and history of glaciation in the North Atlantic region. Nature, 307(5952), 620–623. https://doi.org/10.1038/307620a0 Shen, Z. X., Cao, J. J., Zhang, X. Y., Arimoto, R., Ji, J. F., Balsam, W. L., et al. (2006). Spectroscopic analysis of iron-oxide minerals in aerosol particles from northern China. Science of the Total Environment, 367(2-3), 899–907. https://doi.org/10.1016/j.scitotenv.2006.01.003 Smirnov, A. V., & Tarduno, J. A. (2000). Low-temperature magnetic properties of pelagic sediments (Ocean Drilling Program Site 805C): Tracers of maghemitization and magnetic mineral reduction. Journal of Geophysical Research, 105(B7), 16,457–16,471. https://doi.org/ 10.1029/2000JB900140 Snoeckx, H., Rea, D. K., Jones, C. E., & Ingram, B. L. (1995). Eolian and silica deposition in the central North Pacific: Results from sites 885/886. Proceedings of the Ocean Drilling Program: Scientific Results, 145, 219–230. Spracklen, D. V., Carslaw, K. S., Kulmala, M., Kerminen, V. M., Sihto, S. L., Riipinen, I., et al. (2008). Contribution of particle formation to global cloud condensation nuclei concentrations. Geophysical Research Letters, 35, L06808. https://doi.org/10.1029/2007GL033038 Sugiura, N. (1979). ARM, TRM and magnetic interactions: Concentration dependence. Earth and Planetary Science Letters, 42(3), 451–455. https://doi.org/10.1016/0012-821X(79)90054-2 Sun, J., Zhang, M., & Liu, T. (2001). Spatial and temporal characteristics of dust storms in China and its surrounding regions, 1960–1999: Relations to source area and climate. Journal of Geophysical Research, 106(D10), 10,325–10,333. https://doi.org/10.1029/2000JD900665 Sun, Y., Chen, H., Tada, R., Weiss, D., Lin, M., Toyoda, S., et al. (2013). ESR signal intensity and crystallinity of quartz from Gobi and sandy deserts in East Asia and implication for tracing Asian dust provenance. Geochemistry, Geophysics, Geosystems, 14, 2615–2627. https://doi. org/10.1002/ggge.20162 Sun, Y., Clemens, S. C., An, Z., & Yu, Z. (2006). Astronomical timescale and paleoclimatic implication of stacked 3.6-Myr monsoon records from the Chinese Loess Plateau. Quaternary Science Reviews, 25(1-2), 33–48. https://doi.org/10.1016/j.quascirev.2005.07.005 Sun, Y., & Liu, Q. (2007). Preliminary comparison of eolian depositions in the North Pacific and the Chinese Loess Plateau during the late Pliocene–early Pleistocene. Quaternary Sciences, 27, 263–269. (in Chinese with English abstract) Tanaka, T. Y., & Chiba, M. (2006). A numerical study of the contributions of dust source regions to the global dust budget. Global and Planetary Change, 52(1-4), 88–104. https://doi.org/10.1016/j.gloplacha.2006.02.002 Thompson, R., & Oldfield, F. (1986). Environmental magnetism. London: Allen and Unwin. https://doi.org/10.1007/978-94-011-8036-8 Torrent, J., Liu, Q., Bloemendal, J., & Barrón, V. (2007). Magnetic enhancement and iron oxides in the upper Luochuan loess–paleosol sequence, Chinese Loess Plateau. Soil Science Society of America Journal, 71(5), 1570–1578. https://doi.org/10.2136/sssaj2006.0328 Van Velzen, A. J., & Dekkers, M. J. (1999). Low-temperature oxidation of magnetite in loess-paleosol sequences: A correction of rock magnetic parameters. Studia Geophysica et Geodaetica, 43(4), 357–375. https://doi.org/10.1023/A:1023278901491

ZHANG ET AL. 3375 Journal of Geophysical Research: Solid Earth 10.1002/2017JB014951

Verwey, E. J. W. (1939). Electronic conduction of magnetite (Fe3O4) and its transition point at low temperatures. Nature, 144(3642), 327–328. https://doi.org/10.1038/144327b0 Wang, K., Shi, X., Wu, Y., Zou, J., & Jiang, X. (2014). Characteristics and provenance implications of heavy mineral in core OS03-1 from the east- southern Okhotsk Sea. Acta Oceanologica Sinica, 36, 177–185. (in Chinese with English Abstract) Wang, X., Dong, Z., Zhang, J., & Liu, L. (2004). Modern dust storms in China: An overview. Journal of Arid Environments, 58(4), 559–574. https:// doi.org/10.1016/j.jaridenv.2003.11.009 Weeks, R. J., Roberts, A. P., Verosub, K. L., Okada, M., & Dubuisson, G. J. (1995). Magnetostratigraphy of upper Cenozoic sediments from Leg 145, North Pacific Ocean. Proceedings of the Ocean Drilling Program: Scientific Results, 145, 491–521. Yamazaki, T. (2008). Magnetostatic interactions in deep-sea sediments inferred from first-order reversal curve diagrams: Implications for relative paleointensity normalization. Geochemistry, Geophysics, Geosystems, 9, Q02005. https://doi.org/10.1029/2007GC001797 Yamazaki, T. (2009). Environmental magnetism of Pleistocene sediments in the North Pacific and Ontong-Java Plateau: Temporal variations of detrital and biogenic components. Geochemistry, Geophysics, Geosystems, 10, Q07Z04. https://doi.org/10.1029/2009GC002413 Yamazaki, T., & Ioka, N. (1997). Environmental rock-magnetism of pelagic clay: Implications for Asian eolian input to the North Pacific since the Pliocene. Paleoceanography, 12(1), 111–124. https://doi.org/10.1029/96PA02757 Yamazaki, T., & Shimono, T. (2013). Abundant bacterial magnetite occurrence in oxic red clay. Geology, 41(11), 1191–1194. https://doi.org/ 10.1130/G34782.1 Zhang, D., & Iwasaka, Y. (2004). Size change of Asian dust particles caused by sea salt interaction: Measurements in southwestern Japan. Geophysical Research Letters, 31, L15102. https://doi.org/10.1029/2004GL020087 Zhang, X., Arimoto, R., An, Z., Chen, T., Zhang, G., Zhu, G., & Wang, X. (1993). Atmospheric trace elements over source regions for Chinese dust: Concentrations, sources and atmospheric deposition on the Loess Plateau. Atmospheric Environment. Part A. General Topics, 27(13), 2051–2067. https://doi.org/10.1016/0960-1686(93)90277-6 Zhang, X. Y., Arimoto, R., & An, Z. S. (1997). Dust emission from Chinese desert sources linked to variations in atmospheric circulation. Journal of Geophysical Research, 102(D23), 28,041–28,047. https://doi.org/10.1029/97JD02300 Zhang, Y. G., Ji, J., Balsam, W. L., Liu, L., & Chen, J. (2007). High resolution hematite and goethite records from ODP 1143, South China Sea: Co- evolution of monsoonal precipitation and El Niño over the past 600,000 years. Earth and Planetary Science Letters, 264(1-2), 136–150. https://doi.org/10.1016/j.epsl.2007.09.022

ZHANG ET AL. 3376