US20100064771A1 - Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism - Google Patents

Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism Download PDF

Info

Publication number
US20100064771A1
US20100064771A1 US12/552,276 US55227609A US2010064771A1 US 20100064771 A1 US20100064771 A1 US 20100064771A1 US 55227609 A US55227609 A US 55227609A US 2010064771 A1 US2010064771 A1 US 2010064771A1
Authority
US
United States
Prior art keywords
sno
doped
samples
prepared
temperature
Prior art date
Legal status (The legal status is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the status listed.)
Abandoned
Application number
US12/552,276
Inventor
Alex Punnoose
Current Assignee (The listed assignees may be inaccurate. Google has not performed a legal analysis and makes no representation or warranty as to the accuracy of the list.)
Boise State University
Original Assignee
Boise State University
Priority date (The priority date is an assumption and is not a legal conclusion. Google has not performed a legal analysis and makes no representation as to the accuracy of the date listed.)
Filing date
Publication date
Application filed by Boise State University filed Critical Boise State University
Priority to US12/552,276 priority Critical patent/US20100064771A1/en
Publication of US20100064771A1 publication Critical patent/US20100064771A1/en
Abandoned legal-status Critical Current

Links

Images

Classifications

    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G19/00Compounds of tin
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G19/00Compounds of tin
    • C01G19/02Oxides
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01GCOMPOUNDS CONTAINING METALS NOT COVERED BY SUBCLASSES C01D OR C01F
    • C01G49/00Compounds of iron
    • C01G49/02Oxides; Hydroxides
    • C01G49/06Ferric oxide (Fe2O3)
    • GPHYSICS
    • G01MEASURING; TESTING
    • G01NINVESTIGATING OR ANALYSING MATERIALS BY DETERMINING THEIR CHEMICAL OR PHYSICAL PROPERTIES
    • G01N27/00Investigating or analysing materials by the use of electric, electrochemical, or magnetic means
    • G01N27/72Investigating or analysing materials by the use of electric, electrochemical, or magnetic means by investigating magnetic variables
    • G01N27/74Investigating or analysing materials by the use of electric, electrochemical, or magnetic means by investigating magnetic variables of fluids
    • GPHYSICS
    • G01MEASURING; TESTING
    • G01NINVESTIGATING OR ANALYSING MATERIALS BY DETERMINING THEIR CHEMICAL OR PHYSICAL PROPERTIES
    • G01N33/00Investigating or analysing materials by specific methods not covered by groups G01N1/00 - G01N31/00
    • G01N33/0004Gaseous mixtures, e.g. polluted air
    • G01N33/0009General constructional details of gas analysers, e.g. portable test equipment
    • G01N33/0027General constructional details of gas analysers, e.g. portable test equipment concerning the detector
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01FMAGNETS; INDUCTANCES; TRANSFORMERS; SELECTION OF MATERIALS FOR THEIR MAGNETIC PROPERTIES
    • H01F1/00Magnets or magnetic bodies characterised by the magnetic materials therefor; Selection of materials for their magnetic properties
    • H01F1/01Magnets or magnetic bodies characterised by the magnetic materials therefor; Selection of materials for their magnetic properties of inorganic materials
    • H01F1/40Magnets or magnetic bodies characterised by the magnetic materials therefor; Selection of materials for their magnetic properties of inorganic materials of magnetic semiconductor materials, e.g. CdCr2S4
    • H01F1/401Magnets or magnetic bodies characterised by the magnetic materials therefor; Selection of materials for their magnetic properties of inorganic materials of magnetic semiconductor materials, e.g. CdCr2S4 diluted
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01LSEMICONDUCTOR DEVICES NOT COVERED BY CLASS H10
    • H01L29/00Semiconductor devices adapted for rectifying, amplifying, oscillating or switching, or capacitors or resistors with at least one potential-jump barrier or surface barrier, e.g. PN junction depletion layer or carrier concentration layer; Details of semiconductor bodies or of electrodes thereof  ; Multistep manufacturing processes therefor
    • H01L29/02Semiconductor bodies ; Multistep manufacturing processes therefor
    • H01L29/12Semiconductor bodies ; Multistep manufacturing processes therefor characterised by the materials of which they are formed
    • H01L29/24Semiconductor bodies ; Multistep manufacturing processes therefor characterised by the materials of which they are formed including, apart from doping materials or other impurities, only semiconductor materials not provided for in groups H01L29/16, H01L29/18, H01L29/20, H01L29/22
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01LSEMICONDUCTOR DEVICES NOT COVERED BY CLASS H10
    • H01L29/00Semiconductor devices adapted for rectifying, amplifying, oscillating or switching, or capacitors or resistors with at least one potential-jump barrier or surface barrier, e.g. PN junction depletion layer or carrier concentration layer; Details of semiconductor bodies or of electrodes thereof  ; Multistep manufacturing processes therefor
    • H01L29/02Semiconductor bodies ; Multistep manufacturing processes therefor
    • H01L29/12Semiconductor bodies ; Multistep manufacturing processes therefor characterised by the materials of which they are formed
    • H01L29/26Semiconductor bodies ; Multistep manufacturing processes therefor characterised by the materials of which they are formed including, apart from doping materials or other impurities, elements provided for in two or more of the groups H01L29/16, H01L29/18, H01L29/20, H01L29/22, H01L29/24, e.g. alloys
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2002/00Crystal-structural characteristics
    • C01P2002/70Crystal-structural characteristics defined by measured X-ray, neutron or electron diffraction data
    • C01P2002/72Crystal-structural characteristics defined by measured X-ray, neutron or electron diffraction data by d-values or two theta-values, e.g. as X-ray diagram
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2002/00Crystal-structural characteristics
    • C01P2002/80Crystal-structural characteristics defined by measured data other than those specified in group C01P2002/70
    • C01P2002/82Crystal-structural characteristics defined by measured data other than those specified in group C01P2002/70 by IR- or Raman-data
    • CCHEMISTRY; METALLURGY
    • C01INORGANIC CHEMISTRY
    • C01PINDEXING SCHEME RELATING TO STRUCTURAL AND PHYSICAL ASPECTS OF SOLID INORGANIC COMPOUNDS
    • C01P2006/00Physical properties of inorganic compounds
    • C01P2006/42Magnetic properties
    • HELECTRICITY
    • H01ELECTRIC ELEMENTS
    • H01FMAGNETS; INDUCTANCES; TRANSFORMERS; SELECTION OF MATERIALS FOR THEIR MAGNETIC PROPERTIES
    • H01F10/00Thin magnetic films, e.g. of one-domain structure
    • H01F10/08Thin magnetic films, e.g. of one-domain structure characterised by magnetic layers
    • H01F10/10Thin magnetic films, e.g. of one-domain structure characterised by magnetic layers characterised by the composition
    • H01F10/18Thin magnetic films, e.g. of one-domain structure characterised by magnetic layers characterised by the composition being compounds
    • H01F10/193Magnetic semiconductor compounds
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y10TECHNICAL SUBJECTS COVERED BY FORMER USPC
    • Y10STECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y10S977/00Nanotechnology
    • Y10S977/70Nanostructure
    • Y10S977/773Nanoparticle, i.e. structure having three dimensions of 100 nm or less
    • YGENERAL TAGGING OF NEW TECHNOLOGICAL DEVELOPMENTS; GENERAL TAGGING OF CROSS-SECTIONAL TECHNOLOGIES SPANNING OVER SEVERAL SECTIONS OF THE IPC; TECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y10TECHNICAL SUBJECTS COVERED BY FORMER USPC
    • Y10STECHNICAL SUBJECTS COVERED BY FORMER USPC CROSS-REFERENCE ART COLLECTIONS [XRACs] AND DIGESTS
    • Y10S977/00Nanotechnology
    • Y10S977/70Nanostructure
    • Y10S977/811Of specified metal oxide composition, e.g. conducting or semiconducting compositions such as ITO, ZnOx

Definitions

  • the present invention relates to semiconductors exhibiting room-temperature ferromagnetism.
  • Solid materials are either crystalline or amorphous.
  • a crystalline solid is one in which the atomic arrangement is regularly repeated, and which is likely to exhibit an external morphology of planes making characteristic angles with each other.
  • the mechanical, thermal, optical, electronic, and magnetic properties of crystals are strongly influenced by the periodic arrangement of their atomic cores.
  • a nanoscale particle is a particle having a measurement of 100 nm or less in at least one direction.
  • a crystal may be regarded as a three-dimensional diffraction grating for energetic electromagnetic waves (typically X-rays) of a wavelength comparable with the atomic spacing; the diffraction pattern will provide information about the periodic arrangements of the atoms. Constructive interference of the electromagnetic waves may occur where the following minimum condition (called the Bragg equation) is satisfied:
  • d is the spacing between crystalline planes
  • is the angle of incidence between the beam of X-rays and the parallel crystalline planes
  • n is an integer
  • is the wavelength of the X-rays.
  • a crystalline sample is presented with every ⁇ simultaneously. This is achieved by using a finely powdered crystalline sample in which the crystalline orientations are random. Rays for which one crystallite or another satisfy the Bragg condition emerge from the sample as a series of cones concentric with the incident beam direction. Thus a photographic plate records a series of concentric circles. The spacing and pattern of these circles is used to determine the atomic structure of the crystal.
  • Particle-induced x-ray emission is an analytical technique capable of trace element detection sensitivity of a few parts per million.
  • ions pass through matter, they interact with the electrons in the atoms and occasionally a vacancy is produced by an excited electron.
  • the vacancy is filled by an electron from an outer shell, and an x-ray photon of characteristic energy is emitted.
  • PIXE is 100 times more sensitive than electron micro-analysis systems.
  • Mossbauer spectroscopy is a spectroscopic technique based on the Mossbauer effect.
  • Mossbauer Absorption Spectroscopy a solid sample is exposed to a beam of gamma radiation, and a detector measures the intensity of the beam that is transmitted through the sample.
  • the gamma-ray energy is varied by accelerating the gamma-ray source through a range of velocities with a linear motor. The relative motion between the source and the sample results in an energy shift due to the Doppler effect.
  • gamma-ray intensity is plotted as a function of the source velocity.
  • the sample In x-ray photoelectron spectroscopy, the sample is illuminated with soft x-radiation in an ultrahigh vacuum.
  • the photoelectric effect leads to the production of photoelectrons, the energy spectrum of which can be determined in a beta-ray spectrometer.
  • the difference between the x-ray photon energy, which is known, and the electron energy, which can be measured, results in the binding energy of the orbital from which the electron was expelled. Measurement of the relative areas of the photoelectron peaks allows the composition of the sample to be determined.
  • M is the magnetization
  • ⁇ m is the magnetic susceptibility
  • ⁇ m is the relative permeability
  • is the absolute permeability
  • ⁇ 0 4 ⁇ 10 ⁇ 7 H/m is the permeability of free space.
  • the magnetic response of most solids is dominated by the orientation of permanent dipoles.
  • the response of a magnetic material is usually expressed in terms of either the magnetization M or the magnetic susceptibility ⁇ , where
  • a spinning charged particle constitutes a magnetic dipole.
  • the magnetic dipole moment of an electron is attributed to its “spin,” and creates a magnetic field pointing in a direction perpendicular to the plane in which the electron is spinning, as shown in FIG. 1 .
  • Paramagnetic behavior occurs when the magnetic moments of the various atoms are uncorrelated in the absence of a magnetic field, and the sum total of the magnetic moments tends toward zero. The dipoles do tend to become aligned in a magnetic field.
  • the magnetic susceptibility follows the Curie Law:
  • Antiferromagnetic behavior occurs when the dipoles, or magnetic moments, alternate, causing the sum total of the magnetic moments to tend toward zero. This arrangement is very stable at low temperatures, and the magnetic susceptibility in an applied field is small. When the temperature rises, the efficiency of this dipole-dipole interaction decreases and the magnetic susceptibility increases, until the spins become “free” at the Neel temperature to respond to a field. At even higher temperatures the behavior becomes paramagnetic, and the magnetic susceptibility follows a modified Curie law
  • a ferromagnetic solid, represented in FIG. 2( c ), is ordered with parallel spins below the Curie temperature T c , which results in a spontaneous magnetization M s .
  • the magnitude of this bulk polarization decreases to zero at the Curie temperature T c (which is well below room-temperature in most ferromagnetic solids), and the paramagnetic susceptibility for the disordered spin system at higher temperatures obeys the Curie-Weiss law
  • Ferromagnetism involves the cooperative alignment of permanent atomic dipoles, which arise in atoms having unpaired electrons. The strength of each individual dipole is small, but a completely ordered array of such moments produces a large spontaneous magnetization M s .
  • the low temperature ordering in a ferrimagnetic material is similar to that of an antiferromagnetic material, but the two opposing spin systems have magnetic moments of unequal magnitude, and a net spontaneous magnetization results. This magnetization declines to zero magnitude when the solid is warmed to the Curie point T f , and the behavior is once again paramagnetic at higher temperatures.
  • Hysteresis loops demonstrate a phenomenon wherein a material that did not show any magnetization before the application of a magnetic field exhibits remanent magnetization after the applied magnetic field is removed, as shown in FIG. 3 .
  • the coercivity H c is the value of the magnetic field that must be applied to return the magnetization to zero after the magnetization has been caused to reach its saturation value;
  • the remanence B r (M r in the text below) is the value of the magnetization of the material after the material has been caused to reach its saturation value and then had the applied magnetic field removed.
  • Non-zero values for the coercivity and remanence of a sample imply that the sample is ferromagnetic.
  • Oxide semiconductors have been used to detect gases. However, these have all been non-magnetic oxide semiconductors. Their electrical and semiconducting properties (determined by electrical resistivity, carrier concentration, and carrier mobibility) vary with oxygen stoichiometry. Oxygen stoichiometry can be changed by passing a reducing or oxidizing gas. Thus, traditionally, monitoring the changes in the electrical properties with the type and flow rates of gases has been used as a sensing method.
  • Tin Dioxide SnO 2
  • SnO 2 Tin Dioxide
  • RTFM room-temperature ferromagnetism
  • the present invention is a transition metal-doped semiconductor exhibiting room-temperature ferromagnetism. Preferably, it is manufactured in nanoscale particle form; it also preferably exhibits room-temperature ferromagnetism.
  • the preferred embodiments are iron-doped tin dioxide and cobalt-doped tin dioxide, with the dopant evenly distributed throughout the lattice.
  • FIG. 1 shows a ring of charge rotating about its axis.
  • FIG. 2 shows the low temperature ordering, if any, of neighboring dipoles, and the consequent behavior of spontaneous magnetism and/or susceptibility, for (a), paramagnetism, (b) antiferromagnetism, (c) ferromagnetism, and (d) ferrimagnetism.
  • FIG. 3 shows a schematic magnetization loop for a multi-domain sample of a ferromagnetic solid.
  • H c is the coercivity
  • B r is the remanence of the sample.
  • the dashed curve shows what happens when a nonmagnetized sample is first magnetized.
  • the arrows on the solid curves show the course of a subsequent hysteresis loop.
  • FIG. 4 shows typical PIXE spectra from the Sn 1-x Fe x O 2 samples showing the Fe region. Fe concentrations (x) obtained by PIXE data simulation (solid lines) are given in parentheses.
  • FIGS. 6( a ) and 6 ( b ) show XRD patterns of Sn 1-x Fe x O (prepared at 200° C.), and Sn 1-x Fe x O 2 (prepared at 600° C.), respectively, along with reference lines of orthorhombic SnO 2 (solid lines, marked “O”), romarchite SnO (dotted lines, marked “R”), cassiterite SnO 2 (dashed lines, marked “C”) phases, hematite (marked “H”), and maghemite (marked “M”) phases of Fe 2 O 3 .
  • FIG. 7 shows (a) changes in the lattice parameters a and c of tetragonal SnO calculated using (101) and (110) peaks as a function of Fe percentage, as well as the reported magnitude of the lattice parameters of bulk SnO; (b) changes in the lattice parameters a and c of cassiterite SnO 2 calculated using the (110) and (202) peaks as a function of Fe percentage, as well as the reported magnitude of the lattice parameters of bulk SnO 2 ; (c) particle size of Sn 1-x Fe x O 2 as a function of x calculated from the tetragonal cassiterite XRD peak (110), with particle sizes determined from TEM marked with stars.
  • FIGS. 8( a ) and ( b ) show (a) XRD patterns of 5% Fe-doped samples prepared by annealing the reaction precipitate at different temperatures shown above, along with reference lines of orthorhombic SnO 2 (solid lines, marked “O”), romarchite SnO (dotted lines, marked “R”) cassiterite SnO 2 (dashed lines, marked “C”) phases, hematite (marked “H”) and maghemite (marked “M”) phases of Fe 2 O 3 ; (b) Changes in the lattice parameter a and the lattice volume of cassiterite Sn 0.95 Fe 0.05 O 2 as a function of preparation temperature. The change in the lattice parameter c was minimal.
  • FIGS. 9( a )-( d ) show (a) intensity of the tetragonal cassiterite peak (110) and the orthorhombic fraction x 0 of Sn 1-x Co x O 2 prepared at 600° C. as a function of Co percentage; (b) the particle size of Sn 1-x Co x O 2 as a function of x calculated from the XRD tetragonal cassiterite peak (110). The particle sizes determined from TEM are marked with stars; (c) changes in the lattice parameters a and c of cassiterite SnO 2 as a function of Co percentage.
  • Stars indicate the bulk values of the SnO 2 lattice parameters from XRD reference files; (d) changes in the (110) cassiterite peak position with Co concentration.
  • the lattice parameters were calculated using (110) and (202) peaks of the tetragonal cassiterite phase.
  • FIGS. 10( a ) and ( b ) show diffuse reflectance spectra of Sn 1-x Co x O 2 samples prepared at 600° C.
  • FIGS. 11( a )-( c ) show (a) Raman spectra of Sn 1-x Co x O 2 prepared at 600° C. as a function of x; (b) Raman spectra of pure SnO 2 , 1% Co-doped SnO 2 and a physical mixture of pure SnO 2 and Co 3 O 4 ; (c) the apparent disappearance of SnO 2 Raman peak at 630 cm ⁇ 1 and the emergence of the 617 cm ⁇ 1 peak of Co 3 O 4 for x>0.01.
  • FIGS. 13( a ) and ( b ) show panels (a) and (b) which show TEM images of Sn 0.95 Fe 0.05 O 2 prepared at 350 and 900° C. respectively.
  • FIGS. 14( a )-( c ) show panel (a) which shows the room-temperature M ⁇ umlaut over (R) ⁇ ssbauer spectra of Sn 0.95 Fe 0.05 O.
  • Panels (b) and (c) show the room-temperature M ⁇ umlaut over (R) ⁇ ssbauer spectra of Sn 0.95 Fe 0.05 O 2 prepared at 350 and 600° C., respectively.
  • FIG. 15 shows XPS spectra of Sn 1-x Fe x O 2 prepared at 600° C. with different values of x as indicated. Reference data obtained from maghemite and hematite forms of Fe 2 O 3 prepared under identical synthesis conditions (but with no Sn precursors) are also shown.
  • FIG. 16 is a plot showing the XPS spectra of 5% Fe-doped samples as a function of the annealing temperature. Reference data obtained from maghemite and hematite forms of Fe 2 O 3 prepared at 200 and 600° C., respectively, under identical synthesis conditions (but with no Sn precursors) are also shown.
  • FIG. 17 is a plot showing the XPS spectra of Sn 0.95 Fe 0.05 O 2 prepared at 900° C. along with that obtained from the same sample after removing 10 and 20 nm of surface layer by Ar + ion sputtering.
  • FIGS. 19( a ) and ( b ) show the variation of the atomic percentages of Co, Sn, and O and the Co/Sn ratio as a function of Co concentration calculated using the corresponding XPS peak intensities.
  • FIGS. 20( a )-( c ) show (a) and (c) M vs H data of Sn 1-x Fe x O 2 measured at 300 and 5K, respectively; (b) M- ⁇ p H as a function of H for the Sn 1-x Fe x O 2 samples measured at 300K. Solid lines through the data points in (c) are theoretical fits using the modified Brillouin function for a paramagnetic system.
  • FIG. 22 shows changes in the saturation magnetization M s and remanence Mr, along with the XPS estimate of surface Fe concentration of the Sn 0.95 Fe 0.05 O 2 samples as a function of their preparation temperature.
  • FIGS. 23( a ) and ( b ) show XRD data collected from separate angular regions for Sn 0.95 Fe 0.05 O 2 (data points), pure SnO 2 (dashed line), and pure Fe 2 O 3 , all prepared at 600° C. following identical synthesis procedures. The intensity is plotted on a log scale.
  • FIGS. 24( a )-( c ) show changes in (a) saturation magnetization M s and lattice volume V, (b) remanence M r and the linear paramagnetic component ⁇ p , and (c) interaction parameter T 0 and Curie-Weiss temperature ⁇ (obtained from the paramagnetic component in FIG. 25) of the Sn 1-x Fe x O samples as a function of x.
  • FIG. 30 shows a room temperature hysteresis loop obtained from Sn 0.99 Co 0.01 O 2 prepared at 350° C.
  • the insets (a) and (b) show the hysteresis loops of Sn 0.99 Co 0.01 O 2 prepared at 600° C. and Sn 0.995 Co 0.005 O 2 prepared at 350° C.
  • the lines joining the points are for visual aid.
  • FIGS. 31( a ) and ( b ) show M vs H data for Sn 1-x Fe x O samples measured at 5 and 300K, respectively. Similar M vs H data collected from pure iron oxide (maghemite) prepared under identical conditions (but with no Sn precursors) are also shown. Solid lines through the data points are theoretical fits using the modified Brillouin-function-based form for a paramagnetic system.
  • FIGS. 32( a )-( c ) show (a) the low field region of the room-temperature hysteresis loop of 600° C. prepared Sn 0.99 Co 0.01 O 2 sample showing a coercivity of 9Oe; the inset in (a) shows the complete hysteresis loop of 600° C. prepared Sn 0.99 Co 0.01 O 2 showing saturation of the sample magnetization expected for a ferromagnetic system; (b) variation of the room-temperature coercivity H c and remanence M r of Sn 1-x Co x O 2 prepared at 350° C.
  • FIG. 33 shows M vs H data of Sn 0.95 Co 0.05 O 2 , prepared at the different temperatures indicated, measured at 300K.
  • FIG. 34 shows an embodiment of a gas sensing apparatus using an embodiment of the invented process for detecting a gas.
  • FIG. 35 shows changes in the saturation magnetization of Sn 0.95 Fe 0.05 O 2 versus flow rate of molecular oxygen at 200° C. for a sample prepared at 600° C.
  • the preferred embodiment is a powder comprising an oxide semiconductor that is transition-metal doped; preferably, the semiconductor is SnO 2 . Impurities from other elements could be present and not significantly affect the magnetic properties of the composition.
  • the invention does not have to be in powder form, as the composition manufactured by the preferred process can be converted into other forms, such as films.
  • the transition metal is iron (Fe)
  • the doping concentration is between 0.5% and 10%
  • the Curie temperatures is as high as 850K
  • the composition takes on a powder form with nanoscale particles of which 95% are believed to be less than 100 nm in length
  • there are no phases, or clusters of Fe, in the composition meaning that the Fe is evenly distributed throughout the composition
  • the composition is intrinsically ferromagnetic, meaning that the Fe atoms take the place of the Sn atoms in the lattice, and are substitutionally incorporated into the SnO 2 lattice at the Sn sites.
  • Preparation conditions have a strong effect on the observed magnetic properties and might act as a useful control parameter.
  • Sn 1-x Fe x O 2 is manufactured by the following preferred process, which is less expensive than other methods of manufacturing ferromagnetic oxide semiconductors.
  • Appropriate amounts of tin dichloride (SnCl 2 ) of minimum 99% purity, iron dichloride (FeCl 2 ) of minimum 99.5% purity, and NH 4 OH are added to de-ionized water to produce solutions with molarities of 1, 0.02, and 5M, respectively. All the samples are prepared by reacting the 0.02M FeCl 2 and 1M SnCl 2 solutions at 80° C.
  • iron-doped tin monoxide Sn 1-x Fe x O
  • Pure iron oxide samples have been prepared following identical synthesis procedures without using any SnCl 2 to obtain insight into possible Fe impurity phases that might form under these synthesis conditions.
  • cobalt-doped tin dioxide Sn 1-x Co 2 O 2 , with x being 1% or less, exhibiting room-temperature ferromagnetism
  • magnetic hysteresis loops are observed at 300 K (room-temperature) with coercivity H c ⁇ 630 Oe, saturation magnetization M s ⁇ 0.233 ⁇ B /Co ion and about 31% remenance.
  • the Co is evenly distributed throughout the SnO 2 lattice, and the composite takes on a nanoscale particle form.
  • sol-gel based wet chemistry used to manufacture the two preferred embodiments is preferred over other possible means because it is relatively inexpensive, it intrinsically excludes the segregation of transition metal nanoparticles, it has the ability to kinetically stabilize metastable phases (such as orthorhombic SnO 2 ) and extended solid solutions, and it is very efficient for the controlled syntheses of materials in the nanosize range.
  • metastable phases such as orthorhombic SnO 2
  • the nominal Fe doping concentrations of both Sn 1-x Fe x O 2 and Sn 1-x Fe x O have been confirmed by PIXE measurements.
  • the powder samples were first mixed with a very small amount of polyvinyl alcohol and then palletized using a hand-held press. The samples were then irradiated with a 2.0 MeV He+ ion beam and the x-rays emitted during the de-excitation process within the atoms were analyzed using an x-ray spectrometer.
  • the PIXE data obtained from selected samples of Sn 1-x Fe x O 2 are shown in FIG. 4 .
  • the Fe concentrations shown in Table I estimated by simulating the experimental PIXE spectra after removing the background due to bremsstrahlung, are in reasonable agreement with their nominal concentrations. This confirms that the ratio of Fe to Sn in the resulting product is substantially the same as the ratio of the FeCl 2 and SnCl 2 solutions that were reacted with NH 4 OH.
  • Typical PIXE spectra from Sn 1-x Co x O 2 samples are shown in FIG. 5( a ).
  • the Co concentrations of 1.4, 3.5 and 5.4% estimated by simulating the experimental PIXE spectra after removing the background due to bremsstrahlung are in reasonable, agreement with their nominal concentrations of 1, 3, 5% Co respectively.
  • X-ray diffraction (XRD) studies utilizing the Debye-Scherrer technique were used to determine the crystalline structure of the compositions obtained.
  • the loose powder samples were leveled in the sample holder to ensure a smooth surface and mounted on a fixed horizontal sample plane. Data analyses were carried out using profile fits of selected XRD peaks.
  • the powder Sn 1-x Fe x O 2 samples showed strong XRD peaks due to the cassiterite phase of SnO 2 , with much weaker peaks of the metastable orthorhombic SnO 2 phase.
  • the peak intensities, positions, and widths of the XRD lines changed with x in Sn 1-x Fe x O 2 , as shown in FIG. 6( b ).
  • the XRD patterns of powder Sn 1-x Fe x O samples showed strong peaks of tetragonal SnO with some weak SnO 2 traces, as shown in FIG. 6( a ).
  • the experimentally determined lattice parameters of the pure SnO samples are lower than that reported for pure synthetic bulk romarchite; this may be due to changes in the oxygen stoichiometry and/or particle size effect.
  • Average particle size L of the tetragonal SnO 2 phase was calculated using the width of the (110) peak and the Scherrer relation,
  • FIG. 5( b ) x-ray diffractograms (XRD) of Sn 1-x Co x O 2 samples prepared by annealing the dried precipitate at 600° C. for 6 hours are shown. All samples showed strong peaks due to the rutile-type cassiterite (tetragonal) phase of SnO 2 with relatively weaker peaks of metastable orthorhombic phase. Although cassiterite is the stable phase under ambient conditions, the orthorhombic phase also was observed in SnO 2 synthesized by various methods.
  • XRD x-ray diffractograms
  • FIG. 10( a ) shows a shift of the absorption edge to longer wavelengths/lower energies and a decrease in the band gap of SnO 2 for x ⁇ 0.01.
  • Increasing the Co doping above this level reversed the trend in that the band gap increased gradually and a reduction in the reflectance was observed with Co percentage.
  • the spectra used for the bandgap calculations are plotted in terms of F(R).
  • Raman spectra were collected for the Sn 1-x Co x O 2 samples using a Renishaw S2000 Raman microscope. Samples were all probed using identical instrument conditions: 783 nm diode laser, 1200 line/mm grating, over a Stokes Raman shift range of 50-1000 cm ⁇ 1 . A line focus accessory was also employed, which permitted the collection of photon scatter data from an area 2 ⁇ m by 60 ⁇ m, rather than a discreet 1-2 ⁇ m diameter spot. Incident laser power was not measured; however, power at the laser head was ⁇ 28 mW, which would be expected to produce ⁇ 2-4 mW at the sample. Sample preparation consisted of loosely packing the powder into a stainless steel die accessory, which was then mounted on the microscope stage for probing.
  • FIG. 11( a ) shows the Raman spectra of Sn 1-x Co x O 2 samples as a function of Co concentration.
  • the pure SnO 2 spectrum shown in FIGS. 11( a ) and 11 ( b ) shows the classic cassiterite SnO 2 vibrations at 476 cm ⁇ 1 , 630 cm ⁇ 1 , and 776 cm ⁇ 1 .
  • Addition of 0.5 or 1.0 molar percent Co results in the appearance of two new peaks at 300 cm ⁇ 1 and 692 cm ⁇ 1 . However, no significant change in the SnO 2 peak positions or widths were observed for these doping concentrations.
  • FIG. 11( b ) compares the Raman spectra of undoped SnO 2 and 1% Co doped SnO 2 (both prepared at 600° C.
  • the intensity of the 692 cm ⁇ 1 peak in the 1% doped sample is roughly 1/10 that of the 688 cm ⁇ 1 peak in the physical mixture. This disparity in intensity suggests that even if we assign the 692 cm ⁇ 1 peak to traces of Co 3 O 4 , then there is no more than 0.1% Co 3 O 4 in the 1% Co doped sample.
  • TEM transmission electron microscopy
  • Sn 0.95 Fe 0.05 O 2 particles annealed at 600° C. and 900° C. are shown in FIGS. 13( a ) and 13 ( b ).
  • TEM images also revealed significant differences in the shape of the 600° C. prepared Sn 1-x Co x O 2 particles doped with different percentages of Co.
  • the Sn 0.99 Co 0.01 O 2 nanoparticles shown in FIG. 5( c ) were equiaxed (nearly spherical in shape) with an average diameter of 37 nm, which is about half the average size of the pure SnO 2 particles ( ⁇ 70 nm) prepared under similar conditions.
  • the Sn 0.95 Co 0.05 O 2 particles shown in FIG. 5( d ) were elongated into nanorods with an average aspect ratio of about 3.
  • the Sn 0.99 Co 0.01 O 2 particles had a higher average volume of 25000 nm 3 in comparison to the Sn 0.95 Co 0.05 O 2 particles (average volume ⁇ 10000 nm 3 ). It is estimated that at least 95% of the particles are less than 50 nm long. Electron diffraction patterns taken from an aggregate of particles belonging to each of these samples, shown in the insets of FIGS. 5( c ) and 5 ( d ), revealed characteristic ring patterns that confirm the structure and phase purity measured by XRD. The differences in the particle sizes and shapes observed in the TEM data as well as the systematic changes in the linewidth and intensity of the XRD peaks with increasing Co % seem to be the result of Co incorporation into the SnO 2 lattice. The Co incorporation will most likely result in structural rearrangements to take care of the ionic size differences and charge neutrality. However, more detailed investigations are required to fully understand these structural changes.
  • Mossbauer spectroscopy measurements showed that Sn 0.95 Fe 0.05 O 2 exhibited ferromagnetically ordered Fe 3+ spins when prepared at 350° C., but that these ferromagnetically ordered Fe 3+ spins were converted to a paramagnetic spin system as the preparation temperature increased to 600° C.
  • randomly oriented absorbers were prepared by mixing approximately 30 mg of sample with petroleum jelly in a 0.375 inch thick and 0.5 inch internal diameter Cu holder sealed at one end with clear tape. The holder was entirely filled with the sample mixture and sealed at the other end with tape. Spectra were collected using a 50 mCi (initial strength) 57 Co/Rh source.
  • the velocity transducer MVT-1000 (WissEL) was operated in constant acceleration mode (23 Hz, ⁇ 12 mm/s).
  • An Ar—Kr proportional counter was used to detect the radiation transmitted through the holder, and the counts stored in a multichannel scalar as a function of energy (transducer velocity) using a 1024 channel analyzer. Data were folded to 512 channels to give a flat background and a zero-velocity position corresponding to the center shift (CS or ⁇ ) of a metallic iron foil at room temperature. Calibration spectra were obtained with a 20 ⁇ m thick ⁇ -Fe(m) foil (Amersham, England) placed in exactly the same position as the samples to minimize any errors due to changes in geometry. Sample thickness corrections were not carried out. The data were modeled with RECOIL software (University of Ottawa, Canada) using a Voigt-based spectral fitting routine.
  • FIG. 14( b ) Experimental and fit-derived RT Mossbauer spectra of the Sn 0.95 Fe 0.05 O 2 sample prepared at 350° C. are shown in FIG. 14( b ).
  • the spectrum displayed a well-defined sextet (magnetic component spanning 24% of the spectral area) and a central doublet (paramagnetic component spanning 76% of the spectral area).
  • FIG. 14( b ) shows that the sextet feature (magnetic component of the sample) of the Sn 0.95 Fe 0.05 O 2 prepared at 350° C. is not due to crystalline bulk Fe oxides such as magnetite, hematite, goethite, or maghemite.
  • Magnetite a mixed oxide of Fe 2+ and Fe 3+ , displays two well-defined sextets in its RT Mossbauer spectrum because of the presence of Fe in both tetrahedral (Fe 3+ ) and octahedral sites (Fe 2+ and Fe 3+ at a 1:1 ratio displays a sextet peak due to Fe 2.5+ because of Verwey transition): A and B sites of the inverse spinel structure.
  • the experimental conditions employed to synthesize the binary oxides in this research also implied nonformation of maghemite.
  • the derived Mossbauer parameters of the central doublet which is due to contribution from paramagnetic Fe site(s) to the sample, do not favor the formation of small particle magnetite and goethite.
  • Small-particle Fe oxides such as magnetite ( ⁇ 10 nm), goethite ( ⁇ 15 nm), and hematite ( ⁇ 8 nm) display a doublet at room temperature (well below their magnetic ordering temperature) due to superparamagnetism.
  • the parameters of the doublet in FIG. 14( b ) are inconsistent with superparamagnetic iron oxides.
  • any such iron oxides if present in the superparamagnetic nanoparticle form, cannot produce the hysteresis loops with a finite coercivity ( ⁇ 60 Oe) observed in the magnetic studies. This observation, along with the nonformation of “large” particle magnetite, hematite, and goethite in the sample, implies an absence of conditions that would dictate their formation. Therefore, it is believed that iron oxide does not exist in the preferred embodiment of iron-doped tin dioxide.
  • the Mossbauer data shown in FIG. 14( b ), showing the Mossbauer spectra of room-temperature Sn 0.95 Fe 0.05 O 2 prepared at 350° C. suggest that the sextet results from magnetically ordered Fe 3+ , which constitutes 24% of the incorporated Fe ions.
  • X-ray diffraction (XRD) studies utilizing the Debye-Scherrer technique have also shown that the Fe is evenly distributed throughout the SnO 2 , meaning that there are no phases or clusters of the Fe.
  • the pure iron oxide samples (prepared under identical synthesis conditions, but with no SnCl 2 ) showed maghemite [ ⁇ -Fe 2 O 3 , FIG. 6( a ) and hematite [ ⁇ -Fe 2 O 3 , FIG. 6( b ) phases, with average particle sizes of 22 and 53 nm at 200 and 600° C. preparations, respectively. No trace of iron metal, oxides, or any binary tin-iron phases were observed in any of the doped samples with x ⁇ 0.10.
  • X-ray photoelectron spectroscopy studies showed that Sn 1-x Fe x O 2 and Sn 1-x Co x O 2 prepared according to the methods disclosed herein produce a uniform distribution of the dopant in the entire crystallite.
  • XPS measurements for both Fe-doped and Co-doped SnO 2 were performed on powder samples using a Physical Electronics Quantum 2000 Scanning ESCA Microprobe.
  • the system used a focused monochromatic AlK ⁇ x-ray (1486.7 eV) source and a spherical section analyzer.
  • the instrument had a 16 element multichannel detector.
  • the x-ray beam used was a 105 W, 100 ⁇ m diameter beam that was rastered over a 1.4 mm ⁇ 0.2 mm rectangle on the sample.
  • the powder samples were mounted using a small amount of double-coated carbon conductive tape.
  • the x-ray beam was incident normal to the sample and the photoelectron detector was at 45° off-normal. Data were collected using a pass energy of 46.95 eV. For the Ag 3d 5/2 line, these conditions produce full width at half-maximum of better than 0.98 eV.
  • both the Fe-doped and Co-doped SnO 2 surfaces experienced variable degrees of charging. Low-energy electrons at ⁇ 1 eV, 21 ⁇ A, and low-energy Ar + ions were used to minimize this charging.
  • the BE positions were referenced using the 486.7 eV position for the Sn 3d 5/2 feature for the Sn 1-x Fe x O 2 samples and for the Sn 1-x Co x O 2 samples, and the 486.9 eV position for the Sn 1-x Fe x O samples.
  • XPS spectra were also collected after Ar + ion sputtering using a 4 kV Ar + ion beam rastered over a 4 mm ⁇ 4 mm sample area.
  • the sputter rates were calibrated using a SiO 2 standard with known thickness.
  • FIG. 16 shows the XPS data obtained from the Sn 1-x Fe x O and Sn 1-x Fe x O 2 samples prepared by annealing the same reaction precipitate at 200, 350, 450, 600, 750, and 900° C.
  • the core level Fe peak was observed at ⁇ 56.5 eV and no measurable shifts towards the binding energies expected for magnetite (53.9 eV), hematite (55.7 eV), and maghemite (55.7 eV) were observed when the preparation temperature varied in the 350 to 900° C. range.
  • the Fe doping concentration was 5%, the Fe XPS peaks systematically intensified with increasing preparation temperature.
  • Atomic percentages of Sn, Fe, and O, calculated using the XPS peaks as a function of preparation temperature, are given in Table I. Notwithstanding the difference between the atomic concentrations obtained from PIXE and XPS, the XPS estimates from Sn 0.95 Fe 0.05 O 2 showed a systematic increase in the Fe concentration from 0.7% to 6.2% as the annealing temperature increased from 350 to 900° C. As mentioned above, the lower Fe estimates from the XPS data may be due to the relatively lower detectability of Fe using XPS. In the PIXE measurements discussed above, the Fe concentration of the Sn 1-x Fe x O 2 samples prepared in the entire temperature range was always between 4% and 4.88%, and no systematic variation with preparation temperature was observed.
  • the Co 2p 3/2 and Co 2p 1/2 XPS spectral region of the Sn 1-x Co x O 2 samples are shown in FIG. 18 . Comparing the binding energies of the Co primary and satellite XPS peaks with that observed for Co(0) in Co Metal, Co 2+ in CoO and Co 3+ in ⁇ -Co 2 O 3 , the electronic state of Co in Sn 1-x CO 3 O 2 samples is found to be Co 2+ and that it is not bonded to oxygen as CoO or Co 3 O 4 . It also rules out any metallic Co clusters in the samples, a result well expected for chemically synthesized samples prepared and processed in air.
  • Atomic percentages of Sn, Co, and O calculated using the Sn 3d 5/2 (486.7 eV), O ls (530.65 eV), and Co 2p 3/2 (781.4 eV) peaks are shown in FIG. 19 .
  • three well defined regions are present.
  • 0.01 ⁇ x ⁇ 0.05 range a relatively slower variation is observed suggesting more interstitial incorporation and/or Co 3 O 4 precipitation in agreement with results from XRD, Raman, and optical studies, discussed above.
  • Co ions remove Sn ions from the SnO 2 structure.
  • interstitial incorporation or Co 3 O 4 precipitation there is no Sn removal.
  • changes in the Co/Sn atomic percentage ratio are minimal, indicating lack of further Co incorporation into the SnO 2 lattice.
  • Estimation of the oxygen content in the samples using the O ls (530.65 eV) peak indicated almost stoichiometric Sn/O ratio for the undoped sample.
  • Co doping decreases the oxygen content of the sample, as shown in FIG. 19( a ).
  • the ferromagnetism observed in Sn 1-x Fe x O 2 when prepared in the 350 to 600° C. range, may be due to weak traces of maghemite or magnetite phases of iron oxide formed in the sample.
  • the pure iron oxide samples prepared under identical synthesis conditions showed the formation of pure maghemite when prepared at 200° C. and pure hematite at 600° C.
  • no ferromagnetism was observed in the Sn 1-x Fe x O sample prepared by annealing the precipitate at 200° C., which rules out the presence of any maghemite phase undetected in the XRD data.
  • the M versus H data shown in FIG. 20( c ), obtained from measuring Sn 1-x Fe x O 2 at 5 K, and M versus T data shown in FIG. 21 , obtained from measuring pure iron oxide samples prepared at 200° C. and 600° C. with an applied magnetic field H 500 Oe, ruled out the presence of hematite due to the absence of spin-flop and the strong Morin transitions.
  • the hematite phase is thermally the most stable phase and it undergoes a thermal reduction to the magnetite (Fe 3 O 4 ) phase only above 1200° C.
  • the possible formation of the magnetite phase can also be ruled out. Even if the magnetite phase were formed, the observed disappearance of ferromagnetism when prepared at temperatures above 600° C., as shown in FIG. 22 , would be difficult to understand.
  • the Sn 1-x Fe x O 2 composition showed a strong structure-magnetic property relationship, as shown in FIG. 24( a ), where the increase in the saturation magnetization with Fe concentration matches with the increase in the lattice concentration.
  • Sn 1-x Fe x O showed an expansion of the lattice with increasing Fe concentration, and here no ferromagnetism is observed. Changes in the internal or external lattice volume/pressure have been reported to produce ferromagnetism in itinerant electron metamagnets. Thus, more investigation is required to understand the exact role of structural changes and internal pressure differences in the observed ferromagnetism/paramagnetism of Sn 1-x Fe x O 2 /Sn 1-x Fe x O.
  • the Sn 1-x Fe x O 2 showed ferromagnetic behavior with a Curie temperature of up to 850 K, well above room-temperature, for the 1% Fe-doped sample. All of the Sn 1-x Fe x O 2 samples show well-defined hysteresis loops at 300 K, room-temperature, with remanence M r and saturation magnetization M s increasing gradually with the level of Fe-doping. The ferromagnetic property is stronger when prepared at lower annealing temperatures, and it gradually declines with increasing preparation temperature and eventually disappears completely for preparation temperatures greater than 600° C. In the preferred embodiment, the Sn 1-x Fe x O 2 powder is free of any hematite particles.
  • FIG. 20( a ) The room-temperature M versus H data of Sn 1-x Fe x O 2 , shown in FIG. 20( a ), show a linear component superimposed on a saturating ferromagnetic-like magnetization. If this linear component ⁇ p is subtracted, the M- ⁇ p data show saturation of M expected for a ferromagnetic phase, as shown in FIG. 20( b ).
  • FIG. 20( c ) shows that at 5 K, the ferromagnetic component is overwhelmed by a paramagnetic-like component. Variations of the saturation magnetization M s and ⁇ p obtained from the M versus H data as a function of the Fe-doping are shown in FIGS. 24( a ) and 24 ( b ).
  • T 0 and ⁇ data, shown in FIG. 24( c ), obtained from the M versus H and M versus T data respectively, indicate that the interaction between the disordered (paramagnetic-like) Fe 3+ spins present in Sn 1-x Fe x O 2 is antiferromagnetic (AF) in nature.
  • AF antiferromagnetic
  • M M 0 ⁇ [(2 J+ 1)/2 J]coth[ (2 J+ 1) y/ (2 J )] ⁇ (1/2 J ) coth ( y/ 2 J ) ⁇
  • FIG. 27 X vs. T data of the 1 and 3% Co doped samples are shown along with theoretical fits obtained using the modified Curie-Weiss law
  • the pure hematite form of iron oxide prepared at 600° C. following an identical synthesis procedure but with no Sn precursor, showed a weak magnetization, as shown in FIGS. 20( a ), 20 ( c ), and 21 .
  • the most striking characteristics of bulk hematite include the sharp Morin transition near 263 K in the M versus T data and a spin-flop (SF) transition at H SF ⁇ 67.5 kOe in the M versus H data. Both of these transitions were indeed present in our pure hematite as shown in FIGS. 20( a ) and 21 , albeit with reduced magnitudes which are presumably due to a smaller particle size of ⁇ 53 nm. These transitions were clearly absent in all of the Sn 1-x Fe x O 2 samples, ruling out the presence of any hematite particles.
  • the Sn 1-x Fe x O 2 samples showed well defined hysteresis loops at 300 K, as shown in FIG. 28 , with remanence M r and saturation magnetization M s increasing gradually with the percentage of Fe-doping, as shown in FIGS. 24( a ) and 24 ( b ).
  • FIG. 31( a ) shows the magnetization M of the Sn 1-x Fe x O samples measured at 5 K as a function of applied magnetic field H along with their theoretical estimates obtained using the modified Brillouin-function-based form for a paramagnetic system, given by
  • M M 0 ⁇ [(2 J+ 1)/2 J]coth[ (2 J+ 1) y/ (2 J )] ⁇ (1/2 J ) coth ( y/ 2 J ) ⁇
  • M 0 is the saturation magnetization
  • ⁇ B is the Bohr magneton
  • k is the Boltzman constant.
  • H c the variations of the room-temperature coercivity H c and remanence M r of 350° C.
  • the ferromagnetic regime of Sn 1-x Co x O 2 with x ⁇ 0.01 corresponds to the compositions for which the SnO 2 lattice contracts (see FIGS. 32( b ) and 32 ( c )). This might suggest that the observed ferromagnetism may be related to internal pressure changes. Changes in the internal or external lattice volume/pressure have been reported to produce ferromagnetism in itinerant electron metamagnets.
  • the ferromagnetic component of Sn 1-x Fe x O 2 gradually declines and subsequently disappears as the preparation temperature increases, as shown in FIGS. 14 , 22 , and 33 .
  • Annealing the reaction precipitate at temperature between 350 and 900° C. produces the Sn 1-x Fe x O 2 phase.
  • the M versus H data measured from the Sn 1-x Fe x O 2 prepared by annealing the same reaction precipitate at different temperature, shown in FIG. 33 clearly shows the presence of a ferromagnetic component in samples annealed at 350, 450, and 600° C. Only a linear variation indicating a purely paramagnetic behavior was observed in the sample prepared by annealing at 750 and 900° C.
  • the saturation magnetization M s estimated after subtracting the linear paramagnetic component ⁇ p , is plotted in FIG. 22 . This clearly establishes the fact that the ferromagnetic component is stronger when prepared at lower annealing temperatures and it gradually decreases with increasing preparation temperature, eventually disappearing completely for preparation temperatures >600° C., which is in excellent agreement with the Mossbauer data discussed above.
  • the remanence M r obtained from the hysteresis loops, shown in FIG. 22 also decreases with preparation temperature. This figure also shows that for Sn 1-x Fe x O 2 prepared at or below 600° C., the surface concentration of Fe is less than 4%.
  • the hysteresis loop measured from the 1% Co doped SnO 2 prepared by annealing the dried precipitate at 350° C. is shown.
  • the saturation magnetization is close to that observed for the 600° C. annealed samples, but the loop now exhibits a very large coercivity of ⁇ 630 Oe with good remenance and squareness.
  • FIG. 5( e ) TEM data obtained from the Sn 0.99 Co 0.01 O 2 sample prepared at 350° C.
  • FIG. 5( e ) showed large micrometer sized particles compared to the ⁇ 37nm sized particles, FIG. 5( c ), observed in samples prepared at 600° C. Since the magnetizations of the samples prepared at these two temperatures are comparable, as shown in FIG. 30 , the observed magnetic properties may not be due to nanoscale size effect.
  • the fact that the samples prepared by annealing at 350 and 450° C. showed strong ferromagnetism with excellent reproducibility—as compared to the lower ( ⁇ 20%) reproducibility observed when annealed at 600° C.
  • the inventor has developed the ability to detect a gas by causing the gas to flow across a material and measuring the change in a magnetic property, preferably magnetization, of the material. Changes in the magnetic properties of a material by flowing a gas has never been used as a sensing method.
  • the preferred material for the process is Sn 0.95 Fe 0.05 O 2 , the manufacture and properties of which have been described above. However, other magnetic materials could be used, so long as their magnetic properties change as a gas flows across them.
  • a gas detected using this process is molecular oxygen, O 2 . However, any gas capable of oxidizing or reducing the magnetic material could be used.
  • the preferred apparatus for detecting a gas using this method is shown in FIG. 34 .
  • the gas sensor 10 is made from taking an industry-standard vibrating sample magnetometer (VSM) and adding a gas inlet 20 and mass flow controller (0-300 mL/minute) 22 .
  • the VSM and gas sensor 10 comprise a VSM controller and power supply 12 , a pair of electromagnets 14 of ⁇ 10 kOe, a pair of pickup coils 16 , and a VSM head drive 18 .
  • the gas-sensing material 30 preferably Sn 0.95 Fe 0.05 O 2 , is placed at the end of the vibrating sample rod 19 connected to the VSM head drive.
  • the gas flows out of the gas inlet 20 into a heater (25-600° C.) 21 , then passes by the gas sensing material 30 , and escapes into the atmosphere.
  • the pickup coils 16 measure the magnetization of the gas sensing material 30 ; the magnetization changes as a function of the flow rate.
  • the gas sensor 10 To make the gas sensor 10 usable to detect unknown gases, it must first be calibrated with known gases. The saturation magnetization of Sn 0.95 Fe 0.05 O 2 is shown in FIG. 35 as a function of the flow rate of molecular oxygen. Then, an unknown quantity of gas can flow through the gas sensor 10 , and the measured magnetization compared to a graph similar to that in FIG. 35 to determine the quantity and type of gas flowing through the sensor.
  • the magnetic properties of the gas sensing material 30 change because the carrier-mediated ferromagnetism of Sn 0.95 Fe 0.05 O 2 can be tailored by exposing it to reducing or oxidizing gaseous atmospheres.
  • carrier-mediated ferromagnetism has been developed as a new, efficient gas sensing parameter.
  • a magnetic gas sensor is much more attractive because no electrical contacts are required to detect the response, the detection process requires only a moderate magnetic field to magnetize the sample and a pickup coil to collect the magnetic response of the material, powder samples when used offer a very large surface area and higher sensitivity, magnetic responses are much faster than electrical responses, the lack of electrical contacts and the high magnetic response due to ferromagnetism will further add to the sensitivity of the gas-sensor device, and the operation range can be as high as the Curie temperature T c .
  • the sensing material (before doping), is a surface driven property
  • the gas-sensing and magnetic properties of a doped oxide semiconductor, when used as the material are expected to vary significantly with crystalline size, and therefore depend on the doping concentrations and preparation temperatures.

Abstract

An oxide semiconductor doped with a transition metal and exhibiting room-temperature ferromagnetism is disclosed. The transition metal-doped oxide semiconductor is preferably manufactured in powder form, and the transition metal is preferably evenly distributed throughout the oxide semiconductor. The preferred embodiments are iron-doped tin dioxide and cobalt-doped tin dioxide. Gases may be detected by passing them across a material and measuring the change in magnetic properties of the material; the preferred material is iron-doped tin dioxide.

Description

  • This application is a continuation of U.S. Non-Provisional Application Ser. No. 11/195,573, filed Aug. 1, 2005, which is issuing on Sep. 1, 2009 as U.S. Pat. No. 7,582,222, which claims priority based on U.S. Provisional Application No. 60/598,203, Ferromagnetic Powders of Tin Oxide Nanoparticles Doped with Cobalt and Magnetic Gas Sensor Utilizing Them, filed Jul. 30, 2004, and U.S. Provisional Application No. 60/612,708, Development of High Temperature Ferromagnetism in SnO2 and Paramagnetism in SnO by Fe Doping, filed Sep. 23, 2004, wherein the disclosures of said Non-Provisional Application and of both of said Provisional Applications are hereby incorporated by reference.
  • BACKGROUND OF THE INVENTION
  • 1. Field of the Invention
  • The present invention relates to semiconductors exhibiting room-temperature ferromagnetism.
  • 2. Related Art
  • Solid materials are either crystalline or amorphous. A crystalline solid is one in which the atomic arrangement is regularly repeated, and which is likely to exhibit an external morphology of planes making characteristic angles with each other. In many materials, there are actually a variety of solid phases, each corresponding to a unique crystal structure. These varying crystal phases of the same substance are called “allotropes” or “polymorphs”. The mechanical, thermal, optical, electronic, and magnetic properties of crystals are strongly influenced by the periodic arrangement of their atomic cores. A nanoscale particle is a particle having a measurement of 100 nm or less in at least one direction.
  • A crystal may be regarded as a three-dimensional diffraction grating for energetic electromagnetic waves (typically X-rays) of a wavelength comparable with the atomic spacing; the diffraction pattern will provide information about the periodic arrangements of the atoms. Constructive interference of the electromagnetic waves may occur where the following minimum condition (called the Bragg equation) is satisfied:

  • 2dsin θ=
  • where d is the spacing between crystalline planes, θ is the angle of incidence between the beam of X-rays and the parallel crystalline planes, n is an integer, and λ is the wavelength of the X-rays. This equation will not be satisfied for most angles θ; to investigate crystals of unknown orientation, the rotating crystal method is used, wherein θ is varied as a function of time while X-rays of a single wavelength are presented to a single crystal. A cylinder of photographic film records a spot whenever the Bragg condition is fulfilled.
  • In the Debye-Scherrer technique, instead of using a single X-ray wavelength and a time-dependent angle of incidence, a crystalline sample is presented with every θ simultaneously. This is achieved by using a finely powdered crystalline sample in which the crystalline orientations are random. Rays for which one crystallite or another satisfy the Bragg condition emerge from the sample as a series of cones concentric with the incident beam direction. Thus a photographic plate records a series of concentric circles. The spacing and pattern of these circles is used to determine the atomic structure of the crystal.
  • Particle-induced x-ray emission (PIXE) is an analytical technique capable of trace element detection sensitivity of a few parts per million. When ions pass through matter, they interact with the electrons in the atoms and occasionally a vacancy is produced by an excited electron. When this occurs in an inner shell, the vacancy is filled by an electron from an outer shell, and an x-ray photon of characteristic energy is emitted. By measuring the energy of the photon, one can determine the atomic number of the element and the amount of the element present that can be extracted from the area under the x-ray peak. For identification and quantification of trace elements, PIXE is 100 times more sensitive than electron micro-analysis systems.
  • Mossbauer spectroscopy is a spectroscopic technique based on the Mossbauer effect. In its most common form, Mossbauer Absorption Spectroscopy, a solid sample is exposed to a beam of gamma radiation, and a detector measures the intensity of the beam that is transmitted through the sample. The gamma-ray energy is varied by accelerating the gamma-ray source through a range of velocities with a linear motor. The relative motion between the source and the sample results in an energy shift due to the Doppler effect. In the resulting spectra, gamma-ray intensity is plotted as a function of the source velocity. At velocities responding to the resonant energy levels of the sample, some of the gamma-rays are absorbed, resulting in a dip in the measured intensity and a corresponding dip in the spectrum. The number, positions, and intensities of the dips (also called peaks) provide information about the chemical environment of the absorbing nuclei and can be used to characterize the sample.
  • In x-ray photoelectron spectroscopy, the sample is illuminated with soft x-radiation in an ultrahigh vacuum. The photoelectric effect leads to the production of photoelectrons, the energy spectrum of which can be determined in a beta-ray spectrometer. The difference between the x-ray photon energy, which is known, and the electron energy, which can be measured, results in the binding energy of the orbital from which the electron was expelled. Measurement of the relative areas of the photoelectron peaks allows the composition of the sample to be determined.
  • In discussing magnetic fields, the relationship between the magnetic field intensity H and the corresponding magnetic induction B is
  • B = μ 0 ( H + M ) = μ 0 ( 1 + χ m ) H = μ 0 μ m H = μ H
  • where M is the magnetization, χm is the magnetic susceptibility, μm is the relative permeability, μ is the absolute permeability, and μ0=4π×10−7 H/m is the permeability of free space.
  • The magnetic response of most solids is dominated by the orientation of permanent dipoles. The response of a magnetic material is usually expressed in terms of either the magnetization M or the magnetic susceptibility χ, where

  • M=χH

  • and

  • χ=M/H
  • A spinning charged particle constitutes a magnetic dipole. The magnetic dipole moment of an electron is attributed to its “spin,” and creates a magnetic field pointing in a direction perpendicular to the plane in which the electron is spinning, as shown in FIG. 1.
  • There are four different kinds of magnetic behavior which involve permanent dipoles in a solid, namely paramagnetic, antiferromagnetic, ferromagnetic, and ferrimagnetic. The low temperature ordering, if any, of neighboring dipoles, and the consequent behavior of spontaneous magnetization and/or susceptibility results in hysteresis loops (shown in FIG. 3) in ferromagnetic and ferromagnetic materials.
  • Paramagnetic behavior, shown in FIG. 2( a), occurs when the magnetic moments of the various atoms are uncorrelated in the absence of a magnetic field, and the sum total of the magnetic moments tends toward zero. The dipoles do tend to become aligned in a magnetic field. The magnetic susceptibility follows the Curie Law:

  • χm =C/T
  • where C is the Curie constant of the solid, and T represents the temperature of the solid.
  • Antiferromagnetic behavior, shown in FIG. 2( b), occurs when the dipoles, or magnetic moments, alternate, causing the sum total of the magnetic moments to tend toward zero. This arrangement is very stable at low temperatures, and the magnetic susceptibility in an applied field is small. When the temperature rises, the efficiency of this dipole-dipole interaction decreases and the magnetic susceptibility increases, until the spins become “free” at the Neel temperature to respond to a field. At even higher temperatures the behavior becomes paramagnetic, and the magnetic susceptibility follows a modified Curie law

  • χm =C/(T+θ)
  • A ferromagnetic solid, represented in FIG. 2( c), is ordered with parallel spins below the Curie temperature Tc, which results in a spontaneous magnetization Ms. The magnitude of this bulk polarization decreases to zero at the Curie temperature Tc (which is well below room-temperature in most ferromagnetic solids), and the paramagnetic susceptibility for the disordered spin system at higher temperatures obeys the Curie-Weiss law

  • χm =C/(T−T c)
  • Ferromagnetism involves the cooperative alignment of permanent atomic dipoles, which arise in atoms having unpaired electrons. The strength of each individual dipole is small, but a completely ordered array of such moments produces a large spontaneous magnetization Ms.
  • The low temperature ordering in a ferrimagnetic material, as shown in FIG. 2( d), is similar to that of an antiferromagnetic material, but the two opposing spin systems have magnetic moments of unequal magnitude, and a net spontaneous magnetization results. This magnetization declines to zero magnitude when the solid is warmed to the Curie point Tf, and the behavior is once again paramagnetic at higher temperatures.
  • Hysteresis loops demonstrate a phenomenon wherein a material that did not show any magnetization before the application of a magnetic field exhibits remanent magnetization after the applied magnetic field is removed, as shown in FIG. 3. The coercivity Hc is the value of the magnetic field that must be applied to return the magnetization to zero after the magnetization has been caused to reach its saturation value; the remanence Br (Mr in the text below) is the value of the magnetization of the material after the material has been caused to reach its saturation value and then had the applied magnetic field removed. Non-zero values for the coercivity and remanence of a sample imply that the sample is ferromagnetic.
  • Oxide semiconductors have been used to detect gases. However, these have all been non-magnetic oxide semiconductors. Their electrical and semiconducting properties (determined by electrical resistivity, carrier concentration, and carrier mobibility) vary with oxygen stoichiometry. Oxygen stoichiometry can be changed by passing a reducing or oxidizing gas. Thus, traditionally, monitoring the changes in the electrical properties with the type and flow rates of gases has been used as a sensing method.
  • Tin Dioxide, SnO2, is an oxide semiconductor with a wide band gap of ˜3.6 eV. When prepared with oxygen vacancies, SnO2 becomes an n-type semiconductor. When doped with iron (Fe) or cobalt (Co), SnO2 remains an n-type semiconductor. Development of room-temperature ferromagnetism (RTFM) in conventional semiconductors is currently attracting intense interest due to their potential use in spintronics applications. However, most traditional transition metal-doped magnetic semiconductor systems exhibit ferromagnetism only at temperatures that are well below room temperature.
  • SUMMARY OF THE INVENTION
  • The present invention is a transition metal-doped semiconductor exhibiting room-temperature ferromagnetism. Preferably, it is manufactured in nanoscale particle form; it also preferably exhibits room-temperature ferromagnetism. The preferred embodiments are iron-doped tin dioxide and cobalt-doped tin dioxide, with the dopant evenly distributed throughout the lattice.
  • BRIEF DESCRIPTION OF THE DRAWINGS
  • FIG. 1 shows a ring of charge rotating about its axis.
  • FIG. 2 shows the low temperature ordering, if any, of neighboring dipoles, and the consequent behavior of spontaneous magnetism and/or susceptibility, for (a), paramagnetism, (b) antiferromagnetism, (c) ferromagnetism, and (d) ferrimagnetism.
  • FIG. 3 shows a schematic magnetization loop for a multi-domain sample of a ferromagnetic solid. Hc is the coercivity, and Br is the remanence of the sample. The dashed curve shows what happens when a nonmagnetized sample is first magnetized. The arrows on the solid curves show the course of a subsequent hysteresis loop.
  • FIG. 4 shows typical PIXE spectra from the Sn1-xFexO2 samples showing the Fe region. Fe concentrations (x) obtained by PIXE data simulation (solid lines) are given in parentheses.
  • FIG. 5 shows (a) typical PIXE spectra from the Sn1-xCoxO2 samples showing the Co region, with Co concentrations (x) obtained by data simulation (solid lines) also included; (b) XRD data of Sn1-xCoxO2 powders prepared with different values of x; powder diffraction files of orthorhombic SnO2 (solid thin lines, marked O), tetragonal SnO2 (solid thick lines, marked C), and SnO (thick dashed lines, marked R) phases are also shown; the * indicates weak peaks of Co3O4 observed only in samples with x≧0.08; (c) TEM images of Sn1-xCoxO2 prepared at 600° with x=0.01; the inset shows electron diffraction patterns taken from selected regions of the TEM images; (d) TEM images of Sn1-xCoxO2 prepared at 600° with x=0.05; the inset shows electron diffraction patterns taken from selected regions of the TEM images; (e) TEM images of Sn0.99Co0.01O2 prepared at 350° C.; (f) XRD data of Sn1-xCoxO2 powders prepared with by annealing at different temperatures; the bottom and top panels show data from samples annealed in air and in flowing hydrogen (10% H2 and 90% He) respectively; powder diffraction files of orthorhombic SnO2 (solid thin lines, marked O), tetragonal SnO2 (solid thick lines, marked C), and SnO (thick dashed lines, marked R) phases are also shown.
  • FIGS. 6( a) and 6(b) show XRD patterns of Sn1-xFexO (prepared at 200° C.), and Sn1-xFexO2 (prepared at 600° C.), respectively, along with reference lines of orthorhombic SnO2 (solid lines, marked “O”), romarchite SnO (dotted lines, marked “R”), cassiterite SnO2 (dashed lines, marked “C”) phases, hematite (marked “H”), and maghemite (marked “M”) phases of Fe2O3.
  • FIG. 7 shows (a) changes in the lattice parameters a and c of tetragonal SnO calculated using (101) and (110) peaks as a function of Fe percentage, as well as the reported magnitude of the lattice parameters of bulk SnO; (b) changes in the lattice parameters a and c of cassiterite SnO2 calculated using the (110) and (202) peaks as a function of Fe percentage, as well as the reported magnitude of the lattice parameters of bulk SnO2; (c) particle size of Sn1-xFexO2 as a function of x calculated from the tetragonal cassiterite XRD peak (110), with particle sizes determined from TEM marked with stars.
  • FIGS. 8( a) and (b) show (a) XRD patterns of 5% Fe-doped samples prepared by annealing the reaction precipitate at different temperatures shown above, along with reference lines of orthorhombic SnO2 (solid lines, marked “O”), romarchite SnO (dotted lines, marked “R”) cassiterite SnO2 (dashed lines, marked “C”) phases, hematite (marked “H”) and maghemite (marked “M”) phases of Fe2O3; (b) Changes in the lattice parameter a and the lattice volume of cassiterite Sn0.95Fe0.05O2 as a function of preparation temperature. The change in the lattice parameter c was minimal.
  • FIGS. 9( a)-(d) show (a) intensity of the tetragonal cassiterite peak (110) and the orthorhombic fraction x0 of Sn1-xCoxO2 prepared at 600° C. as a function of Co percentage; (b) the particle size of Sn1-xCoxO2 as a function of x calculated from the XRD tetragonal cassiterite peak (110). The particle sizes determined from TEM are marked with stars; (c) changes in the lattice parameters a and c of cassiterite SnO2 as a function of Co percentage. Stars indicate the bulk values of the SnO2 lattice parameters from XRD reference files; (d) changes in the (110) cassiterite peak position with Co concentration. The lattice parameters were calculated using (110) and (202) peaks of the tetragonal cassiterite phase.
  • FIGS. 10( a) and (b) show diffuse reflectance spectra of Sn1-xCoxO2 samples prepared at 600° C. (a) The changes in the absorption edge with Co concentration; inset in panel (a) shows the changes in the band gap energy estimated from the reflectance data as a function of Co concentration and (b) shows the complete spectra indicating the extent of Co3O4 formation.
  • FIGS. 11( a)-(c) show (a) Raman spectra of Sn1-xCoxO2 prepared at 600° C. as a function of x; (b) Raman spectra of pure SnO2, 1% Co-doped SnO2 and a physical mixture of pure SnO2 and Co3O4; (c) the apparent disappearance of SnO2 Raman peak at 630 cm−1 and the emergence of the 617 cm−1 peak of Co3O4 for x>0.01.
  • FIGS. 12( a)-(c) show panels (a) and (b) which show transmission electron microscopy (TEM) images of Sn1-xFexO2 prepared at 600° C. with x=0.01 and 0.05, respectively. Panel (c) shows the TEM image of Sn1-xFexO2 prepared at 200° C. with x=0.05.
  • FIGS. 13( a) and (b) show panels (a) and (b) which show TEM images of Sn0.95Fe0.05O2 prepared at 350 and 900° C. respectively.
  • FIGS. 14( a)-(c) show panel (a) which shows the room-temperature M{umlaut over (R)}ssbauer spectra of Sn0.95Fe0.05O. Panels (b) and (c) show the room-temperature M{umlaut over (R)}ssbauer spectra of Sn0.95Fe0.05 O 2 prepared at 350 and 600° C., respectively.
  • FIG. 15 shows XPS spectra of Sn1-xFexO2 prepared at 600° C. with different values of x as indicated. Reference data obtained from maghemite and hematite forms of Fe2O3 prepared under identical synthesis conditions (but with no Sn precursors) are also shown.
  • FIG. 16 is a plot showing the XPS spectra of 5% Fe-doped samples as a function of the annealing temperature. Reference data obtained from maghemite and hematite forms of Fe2O3 prepared at 200 and 600° C., respectively, under identical synthesis conditions (but with no Sn precursors) are also shown.
  • FIG. 17 is a plot showing the XPS spectra of Sn0.95Fe0.05O2 prepared at 900° C. along with that obtained from the same sample after removing 10 and 20 nm of surface layer by Ar+ ion sputtering.
  • FIGS. 18( a) and (b) show (a) XPS spectra of Sn1-xCoxO2 prepared at 600° C. as a function of Co percentage and (b) shows similar data of Co3O4 (x=1) and SnO2 (x=0) reference samples prepared under identical synthesis conditions.
  • FIGS. 19( a) and (b) show the variation of the atomic percentages of Co, Sn, and O and the Co/Sn ratio as a function of Co concentration calculated using the corresponding XPS peak intensities.
  • FIGS. 20( a)-(c) show (a) and (c) M vs H data of Sn1-xFexO2 measured at 300 and 5K, respectively; (b) M-χpH as a function of H for the Sn1-xFexO2 samples measured at 300K. Solid lines through the data points in (c) are theoretical fits using the modified Brillouin function for a paramagnetic system.
  • FIG. 21 shows M vs T data measured with H=500 Oe for pure iron oxide samples prepared at 200° C. (maghemite) and 600° C. (hematite).
  • FIG. 22 shows changes in the saturation magnetization Ms and remanence Mr, along with the XPS estimate of surface Fe concentration of the Sn0.95Fe0.05O2 samples as a function of their preparation temperature.
  • FIGS. 23( a) and (b) show XRD data collected from separate angular regions for Sn0.95Fe0.05O2 (data points), pure SnO2 (dashed line), and pure Fe2O3, all prepared at 600° C. following identical synthesis procedures. The intensity is plotted on a log scale.
  • FIGS. 24( a)-(c) show changes in (a) saturation magnetization Ms and lattice volume V, (b) remanence Mr and the linear paramagnetic component χp, and (c) interaction parameter T0 and Curie-Weiss temperature θ (obtained from the paramagnetic component in FIG. 25) of the Sn1-xFexO samples as a function of x.
  • FIG. 25 shows M vs T data measure with H=500 Oe from Sn1-xFexO2 samples. Solid lines are theoretical fits using the modified Curie-Weiss law.
  • FIG. 26 shows M vs T data measured with H=500 Oe from Sn1-xFexO samples. Solid lines are theoretical fits using the modified Curie-Weiss law.
  • FIG. 27 shows temperature variations of the magnetic susceptibility χ, measured with an applied field H=500 Oe, for the 1% and 3% Co doped samples prepared at 600° C. Lines through the data points are theoretical fits using the modified Curie-Weiss law. The inset shows the magnetization of the same samples, measured at 5K as a function of magnetic field. Lines through the data points are theoretical fits using the Brillouin-function-based form for a paramagnetic system.
  • FIG. 28 shows room-temperature hysteresis loops of Sn1-xFexO2 prepared at 600° C. with x=0.05 (main panel) and 0.01 (inset). The lines joining the points are for visual aid.
  • FIG. 29 is a plot of the normalized sample magnetization M of a Sn0.99Fe0.01O2 sample measured with H=10 kOe as a function of temperature, indicating a Curie temperature Tc˜850K.
  • FIG. 30 shows a room temperature hysteresis loop obtained from Sn0.99Co0.01O2 prepared at 350° C. The insets (a) and (b) show the hysteresis loops of Sn0.99Co0.01O2 prepared at 600° C. and Sn0.995Co0.005O2 prepared at 350° C. The lines joining the points are for visual aid.
  • FIGS. 31( a) and (b) show M vs H data for Sn1-xFexO samples measured at 5 and 300K, respectively. Similar M vs H data collected from pure iron oxide (maghemite) prepared under identical conditions (but with no Sn precursors) are also shown. Solid lines through the data points are theoretical fits using the modified Brillouin-function-based form for a paramagnetic system.
  • FIGS. 32( a)-(c) show (a) the low field region of the room-temperature hysteresis loop of 600° C. prepared Sn0.99Co0.01O2 sample showing a coercivity of 9Oe; the inset in (a) shows the complete hysteresis loop of 600° C. prepared Sn0.99Co0.01O2 showing saturation of the sample magnetization expected for a ferromagnetic system; (b) variation of the room-temperature coercivity Hc and remanence Mr of Sn1-xCoxO2 prepared at 350° C. as a function x; inset in (b) shows the expanded view of the low-field region of the hysteresis loop of 350° C. prepared Sn0.995Co0.005O2 sample; and, (c) variation of the saturation magnetization Ms with x of Sn1-xCoxO2 samples prepared at 350 (open circles) and 600° C. (solid stars), and the lattice volume V calculated using the a and c values of FIG. 9( c) (shown with solid squares) as a function of Co-doping concentration for the 600° C. prepared Sn1-xCoxO2.
  • FIG. 33 shows M vs H data of Sn0.95Co0.05O2, prepared at the different temperatures indicated, measured at 300K.
  • FIG. 34 shows an embodiment of a gas sensing apparatus using an embodiment of the invented process for detecting a gas.
  • FIG. 35 shows changes in the saturation magnetization of Sn0.95Fe0.05O2 versus flow rate of molecular oxygen at 200° C. for a sample prepared at 600° C.
  • DESCRIPTION OF THE PREFERRED EMBODIMENTS I. The Materials Used in the Gas Sensing Process
  • The preferred embodiment is a powder comprising an oxide semiconductor that is transition-metal doped; preferably, the semiconductor is SnO2. Impurities from other elements could be present and not significantly affect the magnetic properties of the composition. The invention does not have to be in powder form, as the composition manufactured by the preferred process can be converted into other forms, such as films. In the preferred embodiment, the transition metal is iron (Fe), the doping concentration is between 0.5% and 10%, the Curie temperatures is as high as 850K, the composition takes on a powder form with nanoscale particles of which 95% are believed to be less than 100 nm in length, there are no phases, or clusters of Fe, in the composition, meaning that the Fe is evenly distributed throughout the composition, and the composition is intrinsically ferromagnetic, meaning that the Fe atoms take the place of the Sn atoms in the lattice, and are substitutionally incorporated into the SnO2 lattice at the Sn sites. Preparation conditions have a strong effect on the observed magnetic properties and might act as a useful control parameter.
  • The preferred embodiment, Sn1-xFexO2, is manufactured by the following preferred process, which is less expensive than other methods of manufacturing ferromagnetic oxide semiconductors. Appropriate amounts of tin dichloride (SnCl2) of minimum 99% purity, iron dichloride (FeCl2) of minimum 99.5% purity, and NH4OH are added to de-ionized water to produce solutions with molarities of 1, 0.02, and 5M, respectively. All the samples are prepared by reacting the 0.02M FeCl2 and 1M SnCl2 solutions at 80° C. {molar ratio of x=[Fe]/([Fe]+[Sn])} with a large amount (˜1.5 times the precursor solution volume) of a 5M solution of NH4OH. The resulting precipitate is washed to remove any water-soluble byproducts and annealed in air for three hours at 600° C. to obtain powder samples of Sn1-xFexO2; in the case of the ratios of 0.02 M FeCl2 and 1M SnCl2 in one embodiment, Sn0.98Fe0.02O2 is obtained. Samples of Sn0.95Fe0.05O2 have also been prepared by annealing the same precipitate at temperatures of 350, 450, 750, and 900° C. for the purpose of investigating the effect of the annealing temperature. When the precipitate is annealed at 200° C., iron-doped tin monoxide (Sn1-xFexO) results. Pure iron oxide samples have been prepared following identical synthesis procedures without using any SnCl2 to obtain insight into possible Fe impurity phases that might form under these synthesis conditions.
  • In an alternative embodiment, cobalt-doped tin dioxide, Sn1-xCo2O2, with x being 1% or less, exhibiting room-temperature ferromagnetism, has been developed. In this alternative embodiment, magnetic hysteresis loops are observed at 300 K (room-temperature) with coercivity Hc˜630 Oe, saturation magnetization Ms˜0.233 μB/Co ion and about 31% remenance. SnO2 samples doped with ≦1% Co showed RTFM with significantly high coercivity (˜630 Oe), moderate remenance (˜31%) and better squareness of the hysteresis loop, but had a lower magnetic moment of 0.133 μB/Co ion. However, for x>0.01, this ferromagnetism was completely destroyed and the samples demonstrated paramagnetic behavior. Preferably, the Co is evenly distributed throughout the SnO2 lattice, and the composite takes on a nanoscale particle form.
  • The Sn1-xCoxO2 is preferably prepared using a wet chemical method by reacting 0.02 M CoCl2.6H2O and 1 M SnCl2 with a molar ratio of x=[Co/(Co+Sn)]. A few drops of concentrated HCl are added to ensure dissolution. This solution is added to a 5M solution of NH4OH, and the resulting mixture is heated to 80° C. for several hours. The precipitate is annealed in air at various temperatures for three hours to obtain Sn1-xCoxO2. Chemically synthesized Sn1-xCoxO2 powders have been shown to exhibit RTFM for x≦0.01 when prepared in the 350 to 600° C. range.
  • The sol-gel based wet chemistry used to manufacture the two preferred embodiments is preferred over other possible means because it is relatively inexpensive, it intrinsically excludes the segregation of transition metal nanoparticles, it has the ability to kinetically stabilize metastable phases (such as orthorhombic SnO2) and extended solid solutions, and it is very efficient for the controlled syntheses of materials in the nanosize range.
  • For both Fe- and Co-doped SnO2 with preparation temperatures of <300° C. and >750° C., there is no ferromagnetism and the particles are not nanoscale in size. It is believed that a ferromagnetic powder with nanoscale particles can be achieved with an annealing temperature above 300° C. and below 750° C.
  • A. Confirmation of Nominal Doping Concentrations
  • The nominal Fe doping concentrations of both Sn1-xFexO2 and Sn1-xFexO have been confirmed by PIXE measurements. The powder samples were first mixed with a very small amount of polyvinyl alcohol and then palletized using a hand-held press. The samples were then irradiated with a 2.0 MeV He+ ion beam and the x-rays emitted during the de-excitation process within the atoms were analyzed using an x-ray spectrometer.
  • The PIXE data obtained from selected samples of Sn1-xFexO2 are shown in FIG. 4. The Fe concentrations shown in Table I, estimated by simulating the experimental PIXE spectra after removing the background due to bremsstrahlung, are in reasonable agreement with their nominal concentrations. This confirms that the ratio of Fe to Sn in the resulting product is substantially the same as the ratio of the FeCl2 and SnCl2 solutions that were reacted with NH4OH.
  • TABLE I
    Atomic concentration estimates of Sn1−xFexO and
    Sn1−xFexO2 obtained from PIXE and XPS measurements.
    Fe % Preparation Major XRD Estimated atomic % from
    Nominal from temperature identified Processing XPS
    Fe % PIXE (° C.) phase conditions Fe Sn O
    0 200 Sn1−xFexO As-prepared 29.2 55.4
    1 0.79 200 Sn1−xFexO As-prepared 30.1 61.0
    3 3.07 200 Sn1−xFexO As-prepared
    5 4.54 200 Sn1−xFexO As-prepared 1.2 29.2 61.0
    0 600 Sn1−xFexO2 As-prepared 31.2 61.2
    1 0.66 600 Sn1−xFexO2 As-prepared 31.7 62.2
    3 2.66 600 Sn1−xFexO2 As-prepared
    5 4.888 600 Sn1−xFexO2 As-prepared 3.1 27.7 59.7
    5 350 Sn1−xFexO2 As-prepared 0.7 30.0 60.2
    5 450 Sn1−xFexO2 As-prepared 1.1 31.8 64.2
    5 750 Sn1−xFexO2 As-prepared 3.9 28.6 62.3
    5 4.00 900 Sn1−xFexO2 As-prepared 6.2 23.1 56.3
    5 900 Sn1−xFexO2 10 nm Ar+ ion 4.5 36.3 57.5
    sputtered
    5 900 Sn1−xFexO2 20 nm Ar+ ion 3.4 39.7 56.2
    sputtered
  • Typical PIXE spectra from Sn1-xCoxO2 samples are shown in FIG. 5( a). The Co concentrations of 1.4, 3.5 and 5.4% estimated by simulating the experimental PIXE spectra after removing the background due to bremsstrahlung are in reasonable, agreement with their nominal concentrations of 1, 3, 5% Co respectively.
  • B. Crystalline Structure of the Compositions
  • X-ray diffraction (XRD) studies utilizing the Debye-Scherrer technique were used to determine the crystalline structure of the compositions obtained. XRD spectra were recorded at room temperature on a Phillips X'Pert x-ray diffractometer with a CuKα source (λ=1.5418 Å) is Bragg-Brentano geometry. The loose powder samples were leveled in the sample holder to ensure a smooth surface and mounted on a fixed horizontal sample plane. Data analyses were carried out using profile fits of selected XRD peaks.
  • As shown in FIG. 6( b), the powder Sn1-xFexO2 samples showed strong XRD peaks due to the cassiterite phase of SnO2, with much weaker peaks of the metastable orthorhombic SnO2 phase. The peak intensities, positions, and widths of the XRD lines changed with x in Sn1-xFexO2, as shown in FIG. 6( b). Lattice parameters a and c and the particle size L, as shown in FIGS. 7( b) and 7(c), estimated using the cassiterite (110) and (202) peaks of the nanoscale samples of Sn1-xFexO2, decreased as x increased from 0.005 to 0.05.
  • The XRD patterns of powder Sn1-xFexO samples, on the other hand, showed strong peaks of tetragonal SnO with some weak SnO2 traces, as shown in FIG. 6( a). The lattice parameters a and c, determined using the (101) and (110) peaks, showed an increase with x, as shown in FIG. 7( a). The experimentally determined lattice parameters of the pure SnO samples are lower than that reported for pure synthetic bulk romarchite; this may be due to changes in the oxygen stoichiometry and/or particle size effect.
  • The directly opposite changes in the lattice parameters observed in Sn1-xFexO2 and Sn1-xFexO with Fe doping concentration might reflect the effect of substituting Fe3+ for Sn4+ ions in SnO2 and for Sn2+ ions in SnO. This might require rearrangement of neighboring oxygen ions for charge neutrality.
  • When the 5% Fe-doped Sn1-xFexO2 samples were prepared at different temperatures in the 200 to 900° C. range, the tetragonal SnO phase was observed at 200° C. and showed a gradual conversion to the SnO2 phase with increasing preparation temperature until its apparent disappearance at ≧450° C., as illustrated in FIG. 8( a). The lattice parameter a and the unit cell volume V decreased, and the lattice parameter c increased with increased preparation temperature in the 350 to 600° C. range, as shown in FIG. 8( b). Above 600° C., these trends were reversed, and the lattice volume approached closer to the pure SnO2 range.
  • XRD patterns of the Sn1-xCoxO2 samples showed the formation of tetragonal cassiterite SnO2 with a very small fraction of metastable orthorhombic phase. For x≧0.08, weak peaks of Co3O4 started appearing and gradually strengthened with increasing Co doping, suggesting a saturation limit of Co in SnO2. It is noted that with increasing Co concentration, the intensity of the cassiterite SnO2 phase decreased while the relative concentration of the orthorhombic phase gradually increased. Changes in the XRD peak intensity of the cassiterite phase (Ic110) and the orthorhombic phase fraction (xo) of SnO2 are shown in FIG. 9( a). The orthorhombic fraction xo was calculated using the method of standard additions,
  • x o = K K + ( I c 110 / I o 111 ) ,
  • using K=2.69. Formation of the high-temperature orthorhombic SnO2 phase at ambient conditions has been observed in thin films and nanoscale powders. Nucleation of the metastable orthorhombic phase has been attributed to thin film strains and size-dependent internal pressures due to surface stresses in nanoparticles. Therefore, the increasing orthorhombic fraction of SnO2 with Co concentration indicates that Co doping causes structural disorder and strain, and possible changes in the particle size. The growth of the orthorhombic fraction is fast up to 1% Co, above which a slower growth is observed, as shown in FIG. 9( a). This suggests that the intrinsic doping mechanisms active in the Co concentration regimes above and below 1% may be different.
  • Average particle size L of the tetragonal SnO2 phase was calculated using the width of the (110) peak and the Scherrer relation,
  • L = 0.9 λ B cos θ
  • (where θ is the peak position, λ is the x-ray wavelength and B=(Bm 2−Bs 2)1/2 was estimated using the measured peak width Bm and the instrumental width Bs). These estimates showed that the crystallite size decreased with increased Co doping, as shown in FIG. 9( b). This, combined with the TEM images discuss below, indicates that Co doping inhibits the growth of SnO2 nanoparticles.
  • XRD peak positions showed significant changes with Co doping as shown in FIGS. 9( c) and 9(d). The tetragonal cassiterite SnO2 peaks initially shifted to the higher 2θ angles as x increased to 0.01, as shown in FIG. 9( d). But for x=0.03, there is a dramatic shift to the lower angles followed by moderate changes in the peak positions at higher x. These changes revealed interesting variations in the lattice parameters a and c with Co concentration as shown in FIG. 9( c). The SnO2 lattice parameter a initially decreased due to Co doping for x≦0.01. Such a rapid contraction of the lattice can be understood qualitatively considering the sizes of the ions and their local coordinations. Substitution of 0.69 Å sized Sn4+ ions with 0.58 Å sized Co2+ ions is expected to reduce the interatomic spacing significantly, justifying the initial contraction of the lattice for x≦0.01. The observed rapid expansion of the lattice for x=0.03 indicates a significantly different doping mechanism. At higher doping concentrations, the incorporation of dopant ions in interstitial sites has been reported in some host systems causing somewhat similar structural changes in the lattice parameters.
  • Interstitial incorporation of Co2+ ions might cause significant changes and disorder in the SnO2 structure as well as many dramatic changes in the properties of the material, which is discussed in the following sections. The large difference in the charges and coordination numbers of Sn4+ and Co2+ ions will also contribute to the structural disorder in SnO2 due to the removal of some oxygen ions that were attached to the octahedrally coordinated Sn4+. For x≧0.03, the observed expansion along the a-direction and continued contraction along the c-direction shown in FIG. 9( c) will contribute to changes in the shape of the SnO2 lattice and the nanoparticles. This was indeed observed in the TEM measurements reported below. On increasing the Co concentration to 5%, the spherical SnO2 particles (observed in the undoped as well as 1% Co doped samples) appeared as nanorods with an aspect ratio as high as 3.
  • In FIG. 5( b), x-ray diffractograms (XRD) of Sn1-xCoxO2 samples prepared by annealing the dried precipitate at 600° C. for 6 hours are shown. All samples showed strong peaks due to the rutile-type cassiterite (tetragonal) phase of SnO2 with relatively weaker peaks of metastable orthorhombic phase. Although cassiterite is the stable phase under ambient conditions, the orthorhombic phase also was observed in SnO2 synthesized by various methods. The peak positions of the XRD lines of both phases did not show any measurable change, but the intensities of the peaks of the orthorhombic phase increased while that of the rutile (cassiterite) SnO2 phase decreased with increasing x, as shown in FIG. 5( b). As x increases, the XRD peaks appeared wider, indicating possible changes in the crystallite size and/or strain. No trace of cobalt metal, oxides, or any binary tin cobalt phases, were observed in any of the samples doped with up to 5% Co, which is well above the detection limit (˜1.5%) of the x-ray diffractometer used. At higher x, additional peaks of Co3O4 started appearing. The formation of Sn1-xCoxO2 at high temperatures with x≦0.08 has been reported using a solid state reaction method, although at lower temperatures the solubility was ≦2%. The wet chemical process used in this work is expected to increase the solubility limit.
  • C. Optical Properties of the Cobalt-Doped Tin Dioxide
  • Room-temperature optical spectra in the ultraviolet and visible light wavelength ranges were collected for the Sn1-xCoxO2 samples using a CARY 5000 spectrophotometer fitted with an integrating sphere diffuse reflectance accessory. The spectrophotometer measures reflectance relative to a background scatterer, which was powdered BaSO4. These studies indicated that samples with x≦1 do not have any Co3O4 phases.
  • Preliminary optical characterization of the pure and Co doped SnO2 powders were carried out by measuring the diffuse reflectance at room-temperature. FIG. 10( a) shows a shift of the absorption edge to longer wavelengths/lower energies and a decrease in the band gap of SnO2 for x≦0.01. Increasing the Co doping above this level reversed the trend in that the band gap increased gradually and a reduction in the reflectance was observed with Co percentage. The diffuse reflectance, R, of the sample is related to the Kubelka-Munk function F(R) by the relation F(R)=(1−R)2/2R, where R is the percentage reflectance. The spectra used for the bandgap calculations are plotted in terms of F(R). The bandgap energies of the Sn1-xCoxO2 powders were calculated from their diffuse-reflectance spectra by plotting the square of the Kubelka-Munk function, F(R)2 vs. energy in electron volts. The linear part of the curve was extrapolated to F(R)2=0 to get the direct bandgap energy.
  • Diffuse reflectance measurements carried out on a pure nanoscale Co3O4 reference sample prepared using an identical procedure (with x=1), showed prominent signatures at lower energies as shown in FIG. 10( b). Comparison of the optical spectra of Sn1-xCoxO2 samples with this suggests that samples with x≦0.01 do not have any Co3O4 phase. However, evidence of its presence was observed in all samples with x≧0.03. This indicates that in samples with x=0.03 and 0.05, at least a fraction of the Co atoms precipitate as Co3O4, although XRD shows its formation only for x≧0.08. Since the detection limit of the x-ray diffractometer employed was ˜1.5% (determined from XRD measurements of the physical mixtures of SnO2 (x=0) and Co3O4 (x=1) nanoparticles prepared under identical synthesis conditions), the fraction of Co forming Co3O4 in 3% and 5% Co doped SnO2 should only be below this level. This explains why the lattice parameters, band gap, particle size, shape, and orthorhombic/tetragonal fractions continue to change for x>0.01 although the changes are relatively smaller in this range as illustrated in FIGS. 9 and 10.
  • D. Raman Spectra of the Sn1-xCoxO2 Samples
  • Raman spectra were collected for the Sn1-xCoxO2 samples using a Renishaw S2000 Raman microscope. Samples were all probed using identical instrument conditions: 783 nm diode laser, 1200 line/mm grating, over a Stokes Raman shift range of 50-1000 cm−1. A line focus accessory was also employed, which permitted the collection of photon scatter data from an area 2 μm by 60 μm, rather than a discreet 1-2 μm diameter spot. Incident laser power was not measured; however, power at the laser head was ˜28 mW, which would be expected to produce ˜2-4 mW at the sample. Sample preparation consisted of loosely packing the powder into a stainless steel die accessory, which was then mounted on the microscope stage for probing.
  • FIG. 11( a) shows the Raman spectra of Sn1-xCoxO2 samples as a function of Co concentration. The pure SnO2 spectrum shown in FIGS. 11( a) and 11(b) shows the classic cassiterite SnO2 vibrations at 476 cm−1, 630 cm−1, and 776 cm−1. Addition of 0.5 or 1.0 molar percent Co results in the appearance of two new peaks at 300 cm−1 and 692 cm−1. However, no significant change in the SnO2 peak positions or widths were observed for these doping concentrations. These new Raman peaks at 300 cm−1 and 692 cm−1 may be due to the vibrational modes activated by local structural changes resulting from the substitution of Co2+ ions at the Sn4+ sites. Further addition of Co to 3 or 5 molar percent results in the appearance of peaks at 196 cm−1, 480 cm−1, 520 cm−1, 617 cm−1 and 688 cm−1. These five peaks observed in the samples with x≧0.03 match well with published Raman data of Co3O4. The appearance of Co3O4 vibrational modes is in good agreement with the result from the optical data that predicts at least a fraction of the doped Co precipitates out as Co3O4 for x≧0.03
  • For Sn1-xCoxO2 samples with x≧0.03, there is an apparent disappearance of the SnO2 peaks. This disappearance is most obvious for the 630 cm−1 SnO2 peak as illustrated in FIG. 11( c) and would suggest a loss of the SnO2 phase. However, the 776 cm−1 SnO2 peak is still visible in the 3 and 5% Co doped samples. At best, its intensity has diminished by a factor of 2. The extensive peak broadening and the subsequent disappearance of the 630 cm−1 peak and the loss in intensity of the 776 cm−1 peak are indicative of significant structural modifications and disorder of the SnO2 lattice for x≧0.03. Based on the drastic changes observed in the lattice parameter a, particle size and particle shape, this disappearance of the Raman peak also may be due to the interstitial incorporation of Co2+ ions and the subsequent structural changes.
  • As discussed above, at 0.5 and 1 molar percent of Co, a small Raman peak is present at 692 cm−1. However, when the Co molar percentage is increased to 3, an intense Raman peak at 688 cm−1 appears. The width of the 688 cm−1 peak precludes determination if the 692 cm−1 is still present. Two obvious conclusions are possible: i) the 692 cm−1 peak represents a very small amount of Co3O4, and ii) the 692 cm−1 mode represents a vibrational mode of Sn1-xCoxO2. FIG. 11( b) compares the Raman spectra of undoped SnO2 and 1% Co doped SnO2 (both prepared at 600° C. through identical procedures) and the Raman spectrum of a physical mixture of SnO2 and Co3O4 with 1 molar percent Co (both prepared separately at 600° C.). For the physical mixture, no annealing was performed following the mixing. The Raman spectrum of the physical mixture is clearly a superposition of a SnO2 spectrum and a Co3O4 spectrum. It is also clear that the intensity of the SnO2 peaks in the undoped SnO2, the 1% Co doped SnO2, and the physical mixture samples, are essentially the same. This indicates that the same percentage of SnO2 in the cassiterite form is present in all three samples. The intensity of the 692 cm−1 peak in the 1% doped sample is roughly 1/10 that of the 688 cm−1 peak in the physical mixture. This disparity in intensity suggests that even if we assign the 692 cm−1 peak to traces of Co3O4, then there is no more than 0.1% Co3O4 in the 1% Co doped sample.
  • E. Shape and Size of the Particles in the Compositions
  • The shape and size of the particles in the compositions were determined using transmission electron microscopy. This showed that the Sn1-xFexO2 particles were all elongated with their average aspect rations changing from 1.25 and 70 nm long for x=0.01 to 1.7 and 25 nm long for x=0.05. It is estimated that for ferromagnetic Sn1-xFexO2 in this preferred embodiment, 95% of the particles are shorter than 100 nm long. It is estimated that at least 95% of the Sn1-xCoxO2 particles are less than 50 nm long.
  • High-resolution transmission electron microscopy (TEM) analysis was carried out on a JEOL JEM 2010 microscope with a specified point-to-point resolution of 0.194 nm. The operating voltage of the microscope was 200 kV. All images were digitally recorded with a slow scan CCD camera (image size 1024×1024 pixels), and image processing was carried out using the Digital Micrograph software from Gatan (Pleasant, Calif.). Energy dispersive x-ray spectroscopy (EDX) was carried out using the Oxford Link system attached to the TEM.
  • The transmission electron microscopy measurements showed significant changes in the shape and size of the Sn1-xFexO2 particles depending on the level of Fe-doping and the preparation temperature. Sn1-xFexO2 particles prepared at 600° C. are shown in FIGS. 12( a) and 12(b). These particles were all elongated with their aspect ratios and average length L changing from ˜1.25 and 70 nm, for x=0.01, to 1.7 and 25 nm for x=0.05. Sn0.95Fe0.05O2 particles annealed at 600° C. and 900° C. are shown in FIGS. 13( a) and 13(b). It is estimated that for ferromagnetic Sn1-xFexO2 in this preferred embodiment, 95% of the particles are shorter than 100 nm long. These crystallite sizes match very well with similar estimates obtained from the XRD data, shown in FIG. 7( c). The energy dispersive x-ray spectroscopy measurements carried out on 1% and 5% Fe-doped SnO2 samples showed Fe concentrations in reasonable agreement with the estimates obtained from PIXE studies. The TEM images of the Sn1-xFexO samples showed the presence of large micron-sized particles with a different shape, as shown in FIG. 12( c).
  • TEM images also revealed significant differences in the shape of the 600° C. prepared Sn1-xCoxO2 particles doped with different percentages of Co. The Sn0.99Co0.01O2 nanoparticles shown in FIG. 5( c) were equiaxed (nearly spherical in shape) with an average diameter of 37 nm, which is about half the average size of the pure SnO2 particles (˜70 nm) prepared under similar conditions. In contrast, the Sn0.95Co0.05O2 particles shown in FIG. 5( d) were elongated into nanorods with an average aspect ratio of about 3. The Sn0.99Co0.01O2 particles had a higher average volume of 25000 nm3 in comparison to the Sn0.95Co0.05O2 particles (average volume ˜10000 nm3). It is estimated that at least 95% of the particles are less than 50 nm long. Electron diffraction patterns taken from an aggregate of particles belonging to each of these samples, shown in the insets of FIGS. 5( c) and 5(d), revealed characteristic ring patterns that confirm the structure and phase purity measured by XRD. The differences in the particle sizes and shapes observed in the TEM data as well as the systematic changes in the linewidth and intensity of the XRD peaks with increasing Co % seem to be the result of Co incorporation into the SnO2 lattice. The Co incorporation will most likely result in structural rearrangements to take care of the ionic size differences and charge neutrality. However, more detailed investigations are required to fully understand these structural changes.
  • F. Mossbauer Spectra of Sn1-xFexO2
  • Mossbauer spectroscopy measurements showed that Sn0.95Fe0.05O2 exhibited ferromagnetically ordered Fe3+ spins when prepared at 350° C., but that these ferromagnetically ordered Fe3+ spins were converted to a paramagnetic spin system as the preparation temperature increased to 600° C. For these measurements, randomly oriented absorbers were prepared by mixing approximately 30 mg of sample with petroleum jelly in a 0.375 inch thick and 0.5 inch internal diameter Cu holder sealed at one end with clear tape. The holder was entirely filled with the sample mixture and sealed at the other end with tape. Spectra were collected using a 50 mCi (initial strength) 57Co/Rh source. The velocity transducer MVT-1000 (WissEL) was operated in constant acceleration mode (23 Hz, ±12 mm/s). An Ar—Kr proportional counter was used to detect the radiation transmitted through the holder, and the counts stored in a multichannel scalar as a function of energy (transducer velocity) using a 1024 channel analyzer. Data were folded to 512 channels to give a flat background and a zero-velocity position corresponding to the center shift (CS or δ) of a metallic iron foil at room temperature. Calibration spectra were obtained with a 20 μm thick α-Fe(m) foil (Amersham, England) placed in exactly the same position as the samples to minimize any errors due to changes in geometry. Sample thickness corrections were not carried out. The data were modeled with RECOIL software (University of Ottawa, Canada) using a Voigt-based spectral fitting routine.
  • Three selected samples, Sn0.95Fe0.05O prepared at 200° C., Sn0.95Fe0.05O2 prepared at 350° C., and Sn0.95Fe0.05O2 prepared at 600° C., were investigated using Mossbauer spectroscopy, and their spectra are shown in FIG. 14. The Sn0.95Fe0.05O sample prepared at 200° C. showed a well-defined doublet in FIG. 14( a), suggesting that the incorporated Fe is paramagnetic and in the 3+ oxidation state in an octahedral environment. No evidence of any Fe2+ ions was detected in this sample.
  • Experimental and fit-derived RT Mossbauer spectra of the Sn0.95Fe0.05O2 sample prepared at 350° C. are shown in FIG. 14( b). The spectrum displayed a well-defined sextet (magnetic component spanning 24% of the spectral area) and a central doublet (paramagnetic component spanning 76% of the spectral area). The fit-derived parameters of the Fe sextet are center shift δ=0.38 mm/s, quadrupole shift parameter ε=−0.1 mm/s, and hyperfine magnetic field Bhf=50.3 T, whereas the parameters of the doublet are δ=0.37 mm/s and quadrupole shift ε=−0.1 mm/s.
  • FIG. 14( b) shows that the sextet feature (magnetic component of the sample) of the Sn0.95Fe0.05O2 prepared at 350° C. is not due to crystalline bulk Fe oxides such as magnetite, hematite, goethite, or maghemite. Magnetite, a mixed oxide of Fe2+ and Fe3+, displays two well-defined sextets in its RT Mossbauer spectrum because of the presence of Fe in both tetrahedral (Fe3+) and octahedral sites (Fe2+and Fe3+at a 1:1 ratio displays a sextet peak due to Fe2.5+ because of Verwey transition): A and B sites of the inverse spinel structure. Hematite, on the other hand, displays a well-defined sextet with Bhf=51.8 T, δ=0.37 mm/s, and ε=−0.20 minis, which is not shown in FIG. 14( b). Goethite displays a well-defined sextet with Bhf=38 T, δ=0.37 mm/s, and ε=−0.26 mm/s, which is also not shown in FIG. 14( b). The sextet feature shown in FIG. 14( b) is also unlikely to be due to maghemite, which has Bnf=50 T, δ=0.23 to 0.35 mm/s, and ε=−0.02 mm/s. The experimental conditions employed to synthesize the binary oxides in this research also implied nonformation of maghemite.
  • The derived Mossbauer parameters of the central doublet, which is due to contribution from paramagnetic Fe site(s) to the sample, do not favor the formation of small particle magnetite and goethite. Small-particle Fe oxides such as magnetite (<10 nm), goethite (<15 nm), and hematite (<8 nm) display a doublet at room temperature (well below their magnetic ordering temperature) due to superparamagnetism. The parameters of the doublet in FIG. 14( b), however, are inconsistent with superparamagnetic iron oxides. The derived Mossbauer parameters of magnetite and hematite are δ=0.22 mm/s, and ε=0 0.6 mm/s, and δ=0.35 mm/s, and ε=0.49 mm/s, respectively, while the quadrupole splitting of goethite is around ε=0.48 mm/s. Moreover, any such iron oxides, if present in the superparamagnetic nanoparticle form, cannot produce the hysteresis loops with a finite coercivity (˜60 Oe) observed in the magnetic studies. This observation, along with the nonformation of “large” particle magnetite, hematite, and goethite in the sample, implies an absence of conditions that would dictate their formation. Therefore, it is believed that iron oxide does not exist in the preferred embodiment of iron-doped tin dioxide.
  • Thus, the Mossbauer data shown in FIG. 14( b), showing the Mossbauer spectra of room-temperature Sn0.95Fe0.05O2 prepared at 350° C., suggest that the sextet results from magnetically ordered Fe3+, which constitutes 24% of the incorporated Fe ions. The Mossbauer spectra of room-temperature Sn0.95Fe0.05O2 prepared by annealing the precipitate at 600° C. shown in FIG. 14( c), on the other hand, shows mainly a doublet structure (corresponding to paramagnetic Fe3+) with very weak traces of the sextet lines. This clearly suggests that the ferromagnetically ordered Fe3+ spins are converted to a paramagnetic spin system as the preparation temperature increased from 350 to 600° C.
  • E. The Distribution of the Dopant Throughout the Crystallite
  • 1. Confirmation By X-Ray Diffraction Studies of Sn1-xFexO2
  • X-ray diffraction (XRD) studies utilizing the Debye-Scherrer technique have also shown that the Fe is evenly distributed throughout the SnO2, meaning that there are no phases or clusters of the Fe. The pure iron oxide samples (prepared under identical synthesis conditions, but with no SnCl2) showed maghemite [γ-Fe2O3, FIG. 6( a) and hematite [α-Fe2O3, FIG. 6( b) phases, with average particle sizes of 22 and 53 nm at 200 and 600° C. preparations, respectively. No trace of iron metal, oxides, or any binary tin-iron phases were observed in any of the doped samples with x≦0.10. This is consistent with the reported solubility limit of up to 10% Fe in SnO2 (this, along with the fact that the x-ray diffractometer employed can detect only phases that are ≧1.5%, is why experimental data has been reported only for samples with x≦0.05). The lack of any observable phases shows that the Fe is evenly distributed throughout the SnO2.
  • 2. Confirmation By X-Ray Photoelectron Spectroscopy Studies
  • X-ray photoelectron spectroscopy studies (XPS) showed that Sn1-xFexO2 and Sn1-xCoxO2 prepared according to the methods disclosed herein produce a uniform distribution of the dopant in the entire crystallite. XPS measurements for both Fe-doped and Co-doped SnO2 were performed on powder samples using a Physical Electronics Quantum 2000 Scanning ESCA Microprobe. The system used a focused monochromatic AlKα x-ray (1486.7 eV) source and a spherical section analyzer. The instrument had a 16 element multichannel detector. The x-ray beam used was a 105 W, 100 μm diameter beam that was rastered over a 1.4 mm×0.2 mm rectangle on the sample. The powder samples were mounted using a small amount of double-coated carbon conductive tape. The x-ray beam was incident normal to the sample and the photoelectron detector was at 45° off-normal. Data were collected using a pass energy of 46.95 eV. For the Ag 3d5/2 line, these conditions produce full width at half-maximum of better than 0.98 eV. Although the binding energy (BE) scale was calibrated using the Cu 2p3/2 feature at 932.62±0.05 eV and Au 4f feature at 83.96±0.05 eV for known standards, both the Fe-doped and Co-doped SnO2 surfaces experienced variable degrees of charging. Low-energy electrons at ˜1 eV, 21 μA, and low-energy Ar+ ions were used to minimize this charging. The BE positions were referenced using the 486.7 eV position for the Sn 3d5/2 feature for the Sn1-xFexO2 samples and for the Sn1-xCoxO2 samples, and the 486.9 eV position for the Sn1-xFexO samples. XPS spectra were also collected after Ar+ ion sputtering using a 4 kV Ar+ ion beam rastered over a 4 mm×4 mm sample area. The sputter rates were calibrated using a SiO2 standard with known thickness.
  • The Fe 3p1/2 XPS spectral region of Sn1-xFexO2 (prepared by annealing the precipitate at 600° C.) samples with x=0.01 and 0.05 are shown in FIG. 15. Similar data obtained from hematite and maghemite phases of pure Fe2O3 prepared under identical conditions are also shown for comparison. The XPS peaks are not clearly visible in the 1% Fe-doped samples; however, clear peaks were observed in Sn0.95Fe0.05O2. This discrepancy may be related to the limited Fe detection ability of the XPS system. The XPS peaks of both hematite and maghemite phases occur at ˜55.7 eV, which matches well with literature reports. However, the core level peak of Fe in Sn1-xFexO2, as shown in FIG. 15, and Sn1-xFexO, as shown in FIG. 16, 200° C. data, showed a slight shift to higher binding energies (˜56.5 eV) compared to the Fe oxides, indicating the difference in the atomic environment surrounding the incorporated Fe ions. Fe 3p1/2 has a reported binding energy of 53.9 eV for the magnetite (Fe3O4) form of iron oxide. Thus, the XPS data clearly suggest that the Fe peaks observed from the Sn1-xFexO2 and Sn1-xFexO samples are not arising from any maghemite, hematite, or magnetite inclusions in the samples. The relative peak positions of the Sn and O peaks in the samples did not show any measurable change with Fe concentration, suggesting that their chemical environments did not change significantly. Atomic percentages of Sn, Fe, and O calculated using the Sn 3d5/2, Fe 3p1/2, and O ls peaks are given in Table 1.
  • FIG. 16 shows the XPS data obtained from the Sn1-xFexO and Sn1-xFexO2 samples prepared by annealing the same reaction precipitate at 200, 350, 450, 600, 750, and 900° C. In all these samples, the core level Fe peak was observed at ˜56.5 eV and no measurable shifts towards the binding energies expected for magnetite (53.9 eV), hematite (55.7 eV), and maghemite (55.7 eV) were observed when the preparation temperature varied in the 350 to 900° C. range. Although the Fe doping concentration was 5%, the Fe XPS peaks systematically intensified with increasing preparation temperature. Atomic percentages of Sn, Fe, and O, calculated using the XPS peaks as a function of preparation temperature, are given in Table I. Notwithstanding the difference between the atomic concentrations obtained from PIXE and XPS, the XPS estimates from Sn0.95Fe0.05O2 showed a systematic increase in the Fe concentration from 0.7% to 6.2% as the annealing temperature increased from 350 to 900° C. As mentioned above, the lower Fe estimates from the XPS data may be due to the relatively lower detectability of Fe using XPS. In the PIXE measurements discussed above, the Fe concentration of the Sn1-xFexO2 samples prepared in the entire temperature range was always between 4% and 4.88%, and no systematic variation with preparation temperature was observed. Compared to PIXE, which is responsive to the entire bulk of the sample, XPS is a surface-sensitive technique. Therefore, the increasing differences between the Fe concentrations obtained from these two techniques clearly suggest a gradual and systematic surface diffusion of the doped Fe ions with increasing preparation temperature. This suggests that the Sn1-xFexO2 samples prepared at lower temperature produce a more uniform distribution of Fe in the entire crystallite. On the other hand, samples prepared at higher temperatures showed significant diffusion of the incorporated Fe ions towards the particle surface as preparation temperature increases to 900° C.
  • To further confirm the Fe surface diffusion possibility, XPS spectra were collected from the 900° C. prepared Sn1-xFexO2 sample employing Ar+ ion sputtering to remove surface layers from the powder samples mounted on carbon conductive tape, as shown in FIG. 17. These measurements showed a gradual decrease in the Fe concentration from 6.23% in the as-prepared sample to 3.43% when a 20 nm surface layer is removed by Ar+ ion sputtering (see Table I). This fully supports the above conclusion that the higher XPS estimates of Fe concentration obtained from Sn1-xFexO2 samples prepared at higher temperatures (≧600° C.) are indeed due to Fe surface diffusion.
  • The Co 2p3/2 and Co 2p1/2 XPS spectral region of the Sn1-xCoxO2 samples are shown in FIG. 18. Comparing the binding energies of the Co primary and satellite XPS peaks with that observed for Co(0) in Co Metal, Co2+ in CoO and Co3+ in δ-Co2O3, the electronic state of Co in Sn1-xCO3O2 samples is found to be Co2+ and that it is not bonded to oxygen as CoO or Co3O4. It also rules out any metallic Co clusters in the samples, a result well expected for chemically synthesized samples prepared and processed in air. These results agree well with the 2+ oxidation state of Co with S=3/2 determined from magnetization measurements of paramagnetic samples of Sn1-xCoxO2. The ≧1 eV shift of the Co 2p3/2 peak in the Sn1-xCoxO2 samples compared to that observed from the Co3O4 reference sample suggests that Co is indeed incorporated in the SnO2 lattice and not forming any significant amount of Co oxides. However, no significant change in the Co binding energy is observed with increasing Co doping concentration. Careful analysis of the peak positions of the Sn 3d5/2 (486.7 eV) and O ls (530.65 eV) peaks also did not show any noticeable change in the binding energy with increasing Co concentration.
  • Atomic percentages of Sn, Co, and O calculated using the Sn 3d5/2 (486.7 eV), O ls (530.65 eV), and Co 2p3/2 (781.4 eV) peaks are shown in FIG. 19. In these plots, three well defined regions are present. A rapid increase in the atomic percentage of Co and the Co/Sn ratio, and a decrease in the Sn atomic percentage for x≦0.01, indicate substitutional incorporation of Co at Sn sites. In the 0.01≦x≦0.05 range, a relatively slower variation is observed suggesting more interstitial incorporation and/or Co3O4 precipitation in agreement with results from XRD, Raman, and optical studies, discussed above. In the case of substitutional Co incorporation, Co ions remove Sn ions from the SnO2 structure. However, in the case of interstitial incorporation or Co3O4 precipitation, there is no Sn removal. For x>0.05, changes in the Co/Sn atomic percentage ratio are minimal, indicating lack of further Co incorporation into the SnO2 lattice. Estimation of the oxygen content in the samples using the O ls (530.65 eV) peak indicated almost stoichiometric Sn/O ratio for the undoped sample. Co doping decreases the oxygen content of the sample, as shown in FIG. 19( a). Removal of oxygen atoms from the SnO2 lattice is well expected if Co2+ replaces Sn4+ ions due to charge neutrality requirements. It has been argued that substitutional Co2+ and oxygen vacancies in excess of those necessary for charge neutrality are essential to produce ferromagnetism in Ti1-xCoxO2. A rapid loss in oxygen content, indicated by relatively larger changes in the oxygen atomic percentage, for samples with x≦0.01 (FIG. 19( a)), may be crucial for the ferromagnetism observed only in this narrow Co percentage range of ≦1%.
  • 3. Shown By Absence of Iron Oxide Phases in the Sn1-xFexO2
  • The even distribution of Fe throughout the ferromagnetic Sn1-xFexO2 powder has been shown by the absence of iron oxide phases in the samples. The origin of ferromagnetism in dilute magnetic semiconductor oxides has been extensively studied recently because of the possible presence of weaker secondary phases. This is particularly important when the ferromagnetic component is weak. The fact that the sol-gel preparation of the samples and their subsequent drying and annealing processes were all conducted in air intrinsically eliminates the possibility of forming metallic Fe particles.
  • The possibility was investigated that the ferromagnetism observed in Sn1-xFexO2, when prepared in the 350 to 600° C. range, may be due to weak traces of maghemite or magnetite phases of iron oxide formed in the sample. The pure iron oxide samples prepared under identical synthesis conditions showed the formation of pure maghemite when prepared at 200° C. and pure hematite at 600° C. However, no ferromagnetism was observed in the Sn1-xFexO sample prepared by annealing the precipitate at 200° C., which rules out the presence of any maghemite phase undetected in the XRD data. Therefore, it is unlikely that this phase will appear when the Sn1-xFexO2 sample is prepared by annealing the same precipitate in the 350 to 600° C. range. Investigation of the phase transition of pure iron oxide samples prepared under identical conditions showed that the maghemite phase converted to the hematite phase when annealed at temperatures above 350° C. Thus, it is very unlikely that the maghemite phase of iron oxide is present in the Sn1-xFexO2 samples prepared by annealing at temperatures ≧350° C.
  • The M versus H data shown in FIG. 20( c), obtained from measuring Sn1-xFexO2 at 5 K, and M versus T data shown in FIG. 21, obtained from measuring pure iron oxide samples prepared at 200° C. and 600° C. with an applied magnetic field H=500 Oe, ruled out the presence of hematite due to the absence of spin-flop and the strong Morin transitions. The hematite phase is thermally the most stable phase and it undergoes a thermal reduction to the magnetite (Fe3O4) phase only above 1200° C. Thus, thermodynamically, the possible formation of the magnetite phase can also be ruled out. Even if the magnetite phase were formed, the observed disappearance of ferromagnetism when prepared at temperatures above 600° C., as shown in FIG. 22, would be difficult to understand.
  • Further, careful analysis of the samples using XRD, TEM, and selected area diffractions experiments has ruled out the presence of any iron oxide phases in the Sn1-xFexO2 samples. Finally, the Mossbauer data, XPS spectra, and hysteresis loop parameters obtained from the Sn1-xFexO2 samples clearly ruled out the presence of any bulk or nanoscale magnetite, hematite, maghemite, or goethite phases of iron oxide in the samples.
  • Room-temperature ferromagnetism observed in a Mn-doped ZnO system has been shown to result from a metastable Mn2-xZnxO3-d type phase formed by the diffusion of Zn into Mn oxides. In these studies, peaks due to pure and/or doped manganese oxides were clearly observed in the XRD measurements (plotted on a log scale). In the present work, although the saturation magnetization increases by about four times as Fe concentration increases to 5%, no indication of pure or doped iron oxides or other impurities is observed in the XRD measurements (shown on log scale), as illustrated in FIG. 23. In the selected area electron diffraction, XPS, or Mossbauer spectroscopy studies as well, no evidence for the presence of any such mixed phases in Sn1-xFexO2 in the entire ranges of Fe concentration and preparation temperatures were observed. The sol-gel-based chemical synthesis employed to manufacture the Sn1-xFexO2 is well known to provide a uniform distribution of the dopant in the host system at lower temperatures, as compared to the solid state reaction used in the preparation of Mn-doped ZnO.
  • 4. Incorporation of Fe Into the SnO2 and SnO Lattices
  • The systematic changes in the lattice parameters, particle size, and shape observed in XRD and TEM studies strongly support the progressive incorporation of Fe into the SnO2 and SnO lattices with increasing x. The one-to-one match in the relative changes in the saturation magnetization Ms and lattice volume V, shown in FIG. 24, observed in the Sn1-xFexO2 samples, is very strong evidence against any impurity being the origin of the observed ferromagnetism. The one-to-one match also suggests a strong structure-magnetic property relationship in these samples. The striking agreement between the estimated magnetically ordered Fe3+ spins (˜24%) in the powder samples of Sn0.95Fe0.05O2 and a similar estimate of ferromagnetic Fe3+ spins (˜23%) in pulsed-laser-deposited thin films of Sn0.95Fe0.05O2 further supports the conclusion that the observed ferromagnetism in Sn1-xFexO2 is not due to impurity iron oxide phases formed under the different preparation conditions employed.
  • The conclusion that the Fe is incorporated into the SnO2 and SnO lattices is also supported by the role of the host system. It is well known that the p-type semiconducting behavior SnO results from an excess of oxygen, whereas the existence of oxygen vacancies in SnO2 make it an excellent n-type semiconductor. The XPS data obtained for 1% and 5% Fe-doped SnO showed identical oxygen atomic percentages (see Table I), whereas the oxygen concentration decreased in Sn1-xFexO2 with Fe concentration. The Sn—O distance of 2.057 Å in SnO2 is lower than the 2.223 Å in SnO, and this might influence the overlap of the electron orbitals. Thus, in Sn1-xFexO, Fe doping might favor the formation of antiferromagnetic Fe3+—O2−—Fe3+ groups, whereas Sn1-xFexO2 will have a large number of ferromagnetic Fe3+-[oxygen vacancies]-Fe3+ groups because of the oxygen vacancies. This might explain the observed antiferromagnetic interaction in Sn1-xFexO and ferromagnetism in Sn1-xFexO2.
  • The Sn1-xFexO2 composition showed a strong structure-magnetic property relationship, as shown in FIG. 24( a), where the increase in the saturation magnetization with Fe concentration matches with the increase in the lattice concentration. Sn1-xFexO, on the other hand, showed an expansion of the lattice with increasing Fe concentration, and here no ferromagnetism is observed. Changes in the internal or external lattice volume/pressure have been reported to produce ferromagnetism in itinerant electron metamagnets. Thus, more investigation is required to understand the exact role of structural changes and internal pressure differences in the observed ferromagnetism/paramagnetism of Sn1-xFexO2/Sn1-xFexO.
  • F. Magnetic Properties of the Compositions
  • The Sn1-xFexO2 showed ferromagnetic behavior with a Curie temperature of up to 850 K, well above room-temperature, for the 1% Fe-doped sample. All of the Sn1-xFexO2 samples show well-defined hysteresis loops at 300 K, room-temperature, with remanence Mr and saturation magnetization Ms increasing gradually with the level of Fe-doping. The ferromagnetic property is stronger when prepared at lower annealing temperatures, and it gradually declines with increasing preparation temperature and eventually disappears completely for preparation temperatures greater than 600° C. In the preferred embodiment, the Sn1-xFexO2 powder is free of any hematite particles.
  • Magnetic measurements for both Sn1-xFexO2 and Sn1-xCoxO2 were carried out as a function of temperature (4 to 350 K) and magnetic field (0 to ˜65 kOe) using a commercial magnetometer (Quantum Design, PPMS) equipped with a superconducting magnet. Measurements were carried out on tightly packed powder samples placed in a clear plastic drinking straw. The data reported were corrected for the background signal from the sample holder (clear plastic drinking straw) with diamagnetic susceptibility χ=−4.1×10−8 emu/Oe.
  • 1. Iron Concentration Dependence
  • a. Sn1-xFexO2
  • The room-temperature M versus H data of Sn1-xFexO2, shown in FIG. 20( a), show a linear component superimposed on a saturating ferromagnetic-like magnetization. If this linear component χp is subtracted, the M-χp data show saturation of M expected for a ferromagnetic phase, as shown in FIG. 20( b). FIG. 20( c) shows that at 5 K, the ferromagnetic component is overwhelmed by a paramagnetic-like component. Variations of the saturation magnetization Ms and χp obtained from the M versus H data as a function of the Fe-doping are shown in FIGS. 24( a) and 24(b). These data fit reasonably well with the modified Brillouin function, assuming that J=5/2. This indicates that a fraction of the doped Fe is not participating in the ordered magnetic state, in excellent agreement with the Mossbauer results previously shown in FIG. 14( b). The exact nature of this component becomes more evident from the M versus T data shown in FIG. 25. This showed a paramagnetic variation described by the modified Curie-Weiss law similar to their Sn1-xFexO counterparts as shown in FIG. 26, but offset by an amount χ0; this offset is most likely due to the ferromagnetic component. This also confirms that the linear component χp observed in the room-temperature M versus H data shown in FIG. 20( a) is also due to this paramagnetic contribution present in the sample. T0 and θ data, shown in FIG. 24( c), obtained from the M versus H and M versus T data respectively, indicate that the interaction between the disordered (paramagnetic-like) Fe3+ spins present in Sn1-xFexO2 is antiferromagnetic (AF) in nature.
  • Measurements of the sample magnetization M as a function of magnetic field H and temperature T were carried out using a commercial magnetometer (Quantum Design, PPMS) equipped with a superconducting magnet. The data reported were corrected for the background signal from the sample holder. In the inset of FIG. 27, M vs. H plots of Sn1-xCoxO2 samples (with x=0.01 and 0.03) prepared at 600° C. and measured at 5K are shown along with their theoretical estimates obtained using the Brillouin-function-based form for a paramagnetic system, given by

  • M=M 0{[(2J+1)/2J]coth[(2J+1)y/(2J)]−(1/2J)coth(y/2J)}
  • Where y=(gμBJH)/(kT), M0 is the saturation magnetization, g is the spectroscopic splitting factor (g=2.0023 for free electrons), μB is the Bohr magneton and k is the Boltzmann constant. M vs. H data of the Sn1-xCoxO2 samples fit very well with their theoretical estimates yielding a total angular momentum J=1.81±0.1. These values are in good agreement with experimental magnitudes (˜4.8) reported for paramagnetic Co2+ ions with spin S=3/2 [13].
  • Magnetic susceptibility χ=M/H of the samples measured as a function of temperature at a constant H=500 Oe also showed the expected paramagnetic behavior. In FIG. 27, X vs. T data of the 1 and 3% Co doped samples are shown along with theoretical fits obtained using the modified Curie-Weiss law

  • X=X 0 +[C/(T+θ)]
  • Where X0=1.5(0.2)×10−6 emu/g Oe represents weak non-paramagnetic contribution, Curie constant C=Nμ2/3k is a measure of the paramagnetic ion concentration (N=number of magnetic ions/g, μ=magnetic moment of the ion) and θ is the Curie-Weiss temperature which represents the magnetic exchange interactions between the spins. These fits yield θ=0.18 and 1.55K, and C=0.63×10−4 and 1.6×10−4 emu K/g Oe for x=0.01 and 0.03 respectively.
  • The pure hematite form of iron oxide, prepared at 600° C. following an identical synthesis procedure but with no Sn precursor, showed a weak magnetization, as shown in FIGS. 20( a), 20(c), and 21. The most striking characteristics of bulk hematite include the sharp Morin transition near 263 K in the M versus T data and a spin-flop (SF) transition at HSF˜67.5 kOe in the M versus H data. Both of these transitions were indeed present in our pure hematite as shown in FIGS. 20( a) and 21, albeit with reduced magnitudes which are presumably due to a smaller particle size of ˜53 nm. These transitions were clearly absent in all of the Sn1-xFexO2 samples, ruling out the presence of any hematite particles.
  • The Sn1-xFexO2 samples showed well defined hysteresis loops at 300 K, as shown in FIG. 28, with remanence Mr and saturation magnetization Ms increasing gradually with the percentage of Fe-doping, as shown in FIGS. 24( a) and 24(b). The coercivity Hc was in the range of ˜60 Oe, which is significantly different from the value of Hc=1844 Oe obtained for the pure hematite sample prepared under identical conditions. The existence of a significant coercivity in Sn1-xFexO2 samples clearly rules out their possible origin from nanoscale superparamagnetic particles of iron oxides because when magnetic materials are prepared in nanoscale sizes, they demonstrate superparamagnetic behavior characterized by hysteresis loops with zero coercivity above their blocking temperatures. Absence of bulk iron oxides (or nonsuperparamagnetic particles) was confirmed from the Mossbauer data discussed above. Finally, a Curie temperature Tc=850 K was obtained for the 1% Fe-doped SnO2 sample by measuring M up to 1000 K, as shown in FIG. 29. This Curie temperature is among the highest Curie temperatures reported for transition-metal-doped oxide semiconductors.
  • Some of the samples with 0.5 and 1% Co doping annealed at 600° C. showed a ferromagnetic behavior. In the inset (a) of FIG. 30, a hysteresis loop measured at 300 K from a 1% Co doped SnO2 sample prepared by annealing the precipitate at 600° C. is shown. The magnetization saturates very well for H>3 kOe with a saturation magnitude of 0.133 μB/Co ion. However, the data did not show any significant coercivity or remenance.
  • The systematic growth of both ferromagnetic and paramagnetic contributions in Sn1-xFexO2 with increasing x, as shown in FIGS. 24( a) and 24(b), suggests that the ferromagnetic component is not growing at the expense of the paramagnetic Fe3+ ions as Fe doping increases. Other researchers have proposed a ferromagnetic exchange mechanism involving oxygen vacancies, which form F-centers with trapped electrons, for the observed ferromagnetism in Fe-doped SnO2 thin films. Overlap of the F-center electron orbitals with the d-orbitals of the neighboring Fe3+ spins to form Fe3+-[oxygen vacancies]-Fe3+ groups is crucial for the proposed ferromagnetic coupling. It has been argued that doped Fe3+ spins might also exist as isolated paramagnetic spin systems wherever the F-center mediated ferromagnetic coupling is not achieved due to lack of Fe3+ neighbors and/or oxygen vacancies. In addition, any Fe3+—O2−—Fe3+ superexchange interactions will be antiferromagnetic in nature. As Fe doping concentration increases, both ferromagnetic and paramagnetic/antiferromagnetic components will increase leading to the observed variations shown in Figures (FIGS. 15( a) and 15(b)). It is believed that Sn1-xFexO2 will exhibit ferromagnetism for any value of x up to the solubility limit of Fe in SnO2, or 10%.
  • b. Sn1-xFexO
  • FIG. 31( a) shows the magnetization M of the Sn1-xFexO samples measured at 5 K as a function of applied magnetic field H along with their theoretical estimates obtained using the modified Brillouin-function-based form for a paramagnetic system, given by

  • M=M 0{[(2J+1)/2J]coth[(2J+1)y/(2J)]−(1/2J)coth(y/2J)}
  • where y=gμBJH/k(T+T0), M0 is the saturation magnetization, g=2.0023 is the spectroscopic splitting factor, μB is the Bohr magneton, and k is the Boltzman constant. Based on the Mossbauer data discussed above, this analysis was carried out assuming that J=5/2 (which is expected for Fe3+). Here, T0 is included as a measure of the magnetic interaction between the Fe spins, which prevents complete alignment of the spins even at the highest magnetic fields employed. A larger T0 indicates stronger antiferromagnetic (AF) interactions between the disordered Fe spins. Magnitudes of M0 and T0 obtained from this analysis are shown in Table II. M versus H plots of Sn1-xFexO samples measured at 300 K showed a linear variation owing to the paramagnetic behavior, as shown in FIG. 31( b).
  • TABLE II
    Variations of magnetization M0, interaction temperature T0, Curie constant C,
    and Curie-Weiss temperature θ of Sn1−xFexO as a function of x.
    Sn1−xFexO
    Fe Magnetization Interaction Curie constant Curie-Weiss
    percentage M0 (emu/g) temperature T0 (K) (10−4 emu K/g Oe) Temperature θ (K)
    1 0.25 4.50 0.55 4.17
    5 1.24 3.10 2.81 3.00
  • Magnetization M of the Sn1-xFexO samples measured as a function of temperature T at a constant field H=500 Oe also showed the expected paramagnetic behavior, as shown in FIG. 26, following the modified Curie-Weiss law χ=χ0+C/(T+θ), where χ0=4(3)×10−6 emu/g Oe represents weak nonparamagnetic contributions, Curie constant C=Nμ2/3k (N=number of magnetic ions/g, μ=magnetic moment of the ion), and θ is the Curie-Weiss temperature. These fits showed an increase in C (as well as M0) with x, as shown in Table II, confirming the progressive doping of Fe ions. The positive values of θ indicate AF interactions between the Fe spins as observed in other systems as well. Both θ and T0 decrease with x, as shown in Table II, indicating that the AF interaction decreases with increasing Fe doping. This may suggest that there are competing AF and ferromagnetic interactions as x increases.
  • The pure iron oxide sample prepared under identical conditions as Sn1-xFexO was strongly ferromagnetic, as shown in FIGS. 31( a) and 31(b). M versus T data, shown in FIG. 21, of this sample indicated a blocking temperature TB˜21 K, suggesting the presence of nanoscale ferromagnetic particles. These observations match very well with the XRD data showing the formation of nanoscale maghemite. This also rules out the presence of this phase in the Sn1-xFexO samples, which are all paramagnetic for x≦0.05.
  • 2. Cobalt Concentration Dependence
  • Magnetic measurements carried out on pure SnO2 nanoparticles showed the expected diamagnetism with a negative magnetic susceptibility. Applicant has shown that the Sn1-xCoxO2 samples with x≦0.01 were all ferromagnetic at room-temperature when prepared in the 350 to 600° C. temperature range. FIG. 32( a) shows the room temperature hysteresis loop measured from a Sn0.99Co0.01O2 sample prepared at 600° C. illustrating a clear ferromagnetic behavior with a coercivity Hc=9 Oe. In FIG. 32( b), the variations of the room-temperature coercivity Hc and remanence Mr of 350° C. prepared Sn1-xCoxO2 as a function of Co concentration are shown. The observed coercivities of ˜630 Oe and remanences as high as 31% are among the highest reported for dilute magnetic semiconductors. The observed variation of the saturation magnetization Ms with Co concentration measured from the Sn1-xCoxO2 samples prepared at 350 and 600° C. were comparable as illustrated in FIG. 6 c. However, the coercivities of the Sn0.99Co0.01O2 samples decreased from 630 Oe to 9 Oe as the preparation temperature increased from 350 to 600° C. (see FIGS. 32( a) and 32(b)). This most likely indicates a change in the magnetocrystalline anisotropy due to reasons that are unclear at present. More importantly, irrespective of the preparation temperature, all Sn1-xCoxO2 samples prepared in the 350 to 600° C. showed complete destruction of the ferromagnetism above 1% Co doping and only a paramagnetic behavior was observed in this range, as shown in FIG. 32( b). This disappearance of ferromagnetism cannot be explained by assuming a 1% solubility limit for Co in SnO2 because—(i) the ferromagnetic component due to the soluble part of the doped Co should not be destroyed or overwhelmed by the weak paramagnetic component of the segregated Co3O4 formed for x>0.01, (ii) the orthorhombic SnO2 fraction in the samples, lattice parameters, band gap energy, particle size, and shape of the Sn1-xCoxO2 particles (shown in FIGS. 9 and 10) continued to change with x for x>0.01, indicating >1% Co solubility in SnO2, (iii) the observed disappearance of the Raman peaks (FIG. 11( c)) for x>0.01 is unlikely to happen if additional Co doping is not taking place, and (iv) no evidence of any change in the oxidation state of Co or Sn is observed in the XPS measurements for x≧0.01.
  • The appearance of ferromagnetism in Sn1-xCoxO2 samples with x≦0.01 and its complete absence at higher Co concentrations can be qualitatively understood by comparing the changes in the magnetic and structural properties noticed in the XRD, Raman, and TEM studies of the 600° C. prepared samples. As shown in the previous sections, for x≦0.01 the SnO2 lattice contracts, resulting in the reduction of the distance between nearby Co2+ spins, and possibly triggering a ferromagnetic coupling. Substitution of Sn4+ ions (octahedrally coordinated with six nearest oxygen neighbors) in SnO2 with Co2+ ions will result in the creation of oxygen vacancies and additional charge carriers. It is not clear if this ferromagnetic ordering is carrier mediated or via other mechanisms such as based on localized defects (F-centers). Increasing the Co doping to ≧3% results in a rapid expansion of the SnO2 lattice and significant structural disorder indicated by the rapid broadening and disappearance of the Raman peaks, as shown in FIG. 11( b). Such enormous structural changes might have destroyed the ferromagnetic ordering since the magnetic exchange interaction is extremely sensitive to the distance between the interacting spins.
  • It may be noted that the ferromagnetic regime of Sn1-xCoxO2 with x≦0.01 corresponds to the compositions for which the SnO2 lattice contracts (see FIGS. 32( b) and 32(c)). This might suggest that the observed ferromagnetism may be related to internal pressure changes. Changes in the internal or external lattice volume/pressure have been reported to produce ferromagnetism in itinerant electron metamagnets.
  • 3. Temperature Dependence in Sn1-xFexO2
  • The ferromagnetic component of Sn1-xFexO2 gradually declines and subsequently disappears as the preparation temperature increases, as shown in FIGS. 14, 22, and 33. Annealing the reaction precipitate at temperature between 350 and 900° C. produces the Sn1-xFexO2 phase. The M versus H data measured from the Sn1-xFexO2 prepared by annealing the same reaction precipitate at different temperature, shown in FIG. 33, clearly shows the presence of a ferromagnetic component in samples annealed at 350, 450, and 600° C. Only a linear variation indicating a purely paramagnetic behavior was observed in the sample prepared by annealing at 750 and 900° C. The saturation magnetization Ms, estimated after subtracting the linear paramagnetic component χp, is plotted in FIG. 22. This clearly establishes the fact that the ferromagnetic component is stronger when prepared at lower annealing temperatures and it gradually decreases with increasing preparation temperature, eventually disappearing completely for preparation temperatures >600° C., which is in excellent agreement with the Mossbauer data discussed above. The remanence Mr obtained from the hysteresis loops, shown in FIG. 22, also decreases with preparation temperature. This figure also shows that for Sn1-xFexO2 prepared at or below 600° C., the surface concentration of Fe is less than 4%.
  • Based on the observed changes in the Fe XPS peak intensity shown in FIG. 16 and the comparison of the concentrations estimated from the PIXE and XPS data listed in Table I, it was concluded that the Fe ions diffuse towards the particle surface as the preparation temperature increases. The lattice volume plotted in FIG. 8( b) shows a gradual contraction of the lattice as preparation temperature increases, presumably due to the outward diffusion and rearrangement of the doped Fe ions in the SnO2 lattice as evidenced from the XPS data. However, above 600°, the lattice expands, approaching the undoped SnO2 range, and this may be due to the expulsion of some of the doped Fe ions out of the SnO2 lattice. This suggests that low preparation temperatures provide a relatively uniform distribution of the Fe dopant ions in the host SnO2 lattice, and this favors ferromagnetism. The high surface diffusion of the dopant atoms in samples annealed above 600° C. causes the gradual disappearance of this ferromagnetic behavior.
  • 4. Temperature Dependence in Sn1-xCoxO2
  • To further investigate the role of synthesis parameters on the ferromagnetic behavior of 1% Co doped SnO2, new samples were prepared by annealing the precipitate at temperatures of 250, 350, 450, 600 and 830° C. in air, taking a fresh portion of the dried precipitate each time. The sample annealed at 830° C. showed only the cassiterite phase of SnO2, but those annealed at 600, 450, and 350° C. showed cassiterite and orthorhombic phases, as shown in FIG. 5( f). However, the sample annealed at 250° C. showed very strong peaks due to SnO and much weaker peaks of cassiterite SnO2. This indicates that lower annealing temperatures present a much weaker oxidizing environment, leading to the formation of SnO and presumably oxygen deficient SnO2. Some support for this possibility was obtained from the XRD pattern of a fresh dried precipitate annealed in flowing H2 at 350° C., FIG. 5( f), which looks very similar to the sample prepared in air at 250° C. Based on this result, it is inferred that the lower annealing temperatures reduce the oxygen stoichiometry of SnO2 and eventually converts to SnO at annealing temperatures <350° C.
  • In the main panel of FIG. 30, the hysteresis loop measured from the 1% Co doped SnO2 prepared by annealing the dried precipitate at 350° C. is shown. The saturation magnetization is close to that observed for the 600° C. annealed samples, but the loop now exhibits a very large coercivity of ˜630 Oe with good remenance and squareness. These observations were verified on several batches of samples prepared under identical conditions. Such large coercivities and squareness of the loop are seldom observed in DMS's at room temperature. Similar reproducible hysteresis loops were also observed in samples annealed at 450° C. However, no substantial ferromagnetic behavior was detected in samples annealed in air at 250° C. or in H2 at 350° C. presumably due to extensive oxygen loss and transformation to the SnO phase. TEM data obtained from the Sn0.99Co0.01O2 sample prepared at 350° C., FIG. 5( e), showed large micrometer sized particles compared to the ˜37nm sized particles, FIG. 5( c), observed in samples prepared at 600° C. Since the magnetizations of the samples prepared at these two temperatures are comparable, as shown in FIG. 30, the observed magnetic properties may not be due to nanoscale size effect. The fact that the samples prepared by annealing at 350 and 450° C. showed strong ferromagnetism with excellent reproducibility—as compared to the lower (˜20%) reproducibility observed when annealed at 600° C. and its complete absence in samples annealed at 830° C.—may be closely linked to the differences in oxygen stoichiometry of these samples. Theoretically, carrier-mediated ferromagnetism in n-type oxide semiconductors is strongly related to the concentration of oxygen vacancies.
  • In conclusion, it has been shown that powder samples of chemically synthesized Sn1-xCoxO2 powders with x≦0.01 exhibit RTFM. These samples exhibit significantly high coercivity (˜630 Oe) and good squareness of the loop, but with low magnetic moment of 0.133 μB/Co ion. Based on the XRD, PIXE, TEM and magnetic data, the observed ferromagnetic interactions seem to be controlled by the oxygen stoichiometry.
  • II. The Gas Sensing Process
  • The inventor has developed the ability to detect a gas by causing the gas to flow across a material and measuring the change in a magnetic property, preferably magnetization, of the material. Changes in the magnetic properties of a material by flowing a gas has never been used as a sensing method. The preferred material for the process is Sn0.95Fe0.05O2, the manufacture and properties of which have been described above. However, other magnetic materials could be used, so long as their magnetic properties change as a gas flows across them. A gas detected using this process is molecular oxygen, O2. However, any gas capable of oxidizing or reducing the magnetic material could be used. The preferred apparatus for detecting a gas using this method is shown in FIG. 34. The gas sensor 10 is made from taking an industry-standard vibrating sample magnetometer (VSM) and adding a gas inlet 20 and mass flow controller (0-300 mL/minute) 22. The VSM and gas sensor 10 comprise a VSM controller and power supply 12, a pair of electromagnets 14 of ±10 kOe, a pair of pickup coils 16, and a VSM head drive 18. The gas-sensing material 30, preferably Sn0.95Fe0.05O2, is placed at the end of the vibrating sample rod 19 connected to the VSM head drive. The gas flows out of the gas inlet 20 into a heater (25-600° C.) 21, then passes by the gas sensing material 30, and escapes into the atmosphere. The pickup coils 16 measure the magnetization of the gas sensing material 30; the magnetization changes as a function of the flow rate.
  • To make the gas sensor 10 usable to detect unknown gases, it must first be calibrated with known gases. The saturation magnetization of Sn0.95Fe0.05O2 is shown in FIG. 35 as a function of the flow rate of molecular oxygen. Then, an unknown quantity of gas can flow through the gas sensor 10, and the measured magnetization compared to a graph similar to that in FIG. 35 to determine the quantity and type of gas flowing through the sensor.
  • The magnetic properties of the gas sensing material 30 change because the carrier-mediated ferromagnetism of Sn0.95Fe0.05O2 can be tailored by exposing it to reducing or oxidizing gaseous atmospheres. Thus, carrier-mediated ferromagnetism has been developed as a new, efficient gas sensing parameter.
  • Compared to their semiconductor gas sensor counterparts which measure changes in electrical properties, a magnetic gas sensor is much more attractive because no electrical contacts are required to detect the response, the detection process requires only a moderate magnetic field to magnetize the sample and a pickup coil to collect the magnetic response of the material, powder samples when used offer a very large surface area and higher sensitivity, magnetic responses are much faster than electrical responses, the lack of electrical contacts and the high magnetic response due to ferromagnetism will further add to the sensitivity of the gas-sensor device, and the operation range can be as high as the Curie temperature Tc.
  • Since the oxygen stoichiometry in SnO2, the sensing material (before doping), is a surface driven property, the gas-sensing and magnetic properties of a doped oxide semiconductor, when used as the material, are expected to vary significantly with crystalline size, and therefore depend on the doping concentrations and preparation temperatures.
  • Although this invention has been described above with reference to particular means, materials and embodiments, it is to be understood that the invention is not limited to these disclosed particulars, but extends instead to all equivalents within the scope of the following claims.

Claims (23)

1. An oxide semiconductor material comprising Fe-doped tin oxide nanoparticles, wherein the Fe-doped tin oxide nanoparticles exhibit room-temperature ferromagnetism.
2. The material of claim 1 wherein the Fe-doped tin oxide nanoparticles are Fe-doped SnO2 nanoparticles wherein the Fe is evenly distributed through the SnO2 lattice and the material of comprises no iron oxide phases.
3. The material of claim 1 wherein the Fe-doped tin oxide nanoparticles are Fe-doped SnO2 nanoparticles wherein the Fe is evenly distributed through the SnO2 lattice and the material of comprises no iron oxide phases of magnetite, hematite, maghemite, or goethite.
4. The material of claim 1 wherein no traces of iron, iron oxides, or tin metal are observable by X-ray diffraction studies utilizing the Debye-Scherrer technique.
5. The material of claim 1 having a Curie temperature of at least 850 K.
6. The material of claim 1 having a coercivity of at least 60 Oe.
7. The material of claim 1 wherein the Fe-doped tin oxide nanoparticles are Sn0.95Fe0.05O2.
8. Oxide semiconductor powder comprising Fe-doped SnO2 nanoparticles made by the process comprising:
adding SnCl2, FeCl2, and NH4OH to water to produce at least one solution;
reacting the at least one solution together to produce a precipitate; and
annealing the precipitate.
9. The powder of claim 8 wherein the ratio of SnCl2 to FeCl2 is between 200 to 1 and 20 to 1.
10. The powder of claim 8 wherein the precipitate is annealed at a temperature between 350° C. and 600° C.
11. The powder of claim 10 wherein the ratio of SnCl2 to FeCl2 is between 200 to 1 and 20 to 1.
12. An oxide semiconductor material comprising Fe-doped tin oxide exhibiting room-temperature ferromagnetism, wherein the Fe atoms take the place of Sn atoms in the SnO2 lattice so that the Fe atoms are substitutionally incorporated into the SnO2 lattice at the Sn sites, and wherein the oxide semiconductor material comprises no iron oxide phases.
13. The oxide semiconductor material of claim 12, wherein said material is a film comprising said Fe-doped tin oxide exhibiting room-temperature ferromagnetism.
14. The oxide semiconductor material of claim 13, wherein said film is a thin film of Sn0.95Fe0.05O2.
15. An oxide semiconductor material comprising Fe-doped SnO2 particles, wherein the Fe-doped SnO2 particles exhibit room-temperature ferromagnetism and 95% of which particles are less than 100 nm in length.
16. A method of detecting a gas comprising:
causing the gas to flow across a material that comprises transition-metal-doped tin oxide that exhibits room-temperature ferromagnetism; and
measuring a change in a magnetic property of the material.
17. The method of claim 16 wherein the transition-metal-doped tin oxide is Fe-doped SnO2 nanoparticles, wherein Fe takes the place of Sn atoms in the SnO2 lattice so that the Fe atoms are substitutionally incorporated into the SnO2 lattice at the Sn sites.
18. The method of claim 17 wherein the nanoparticles comprise no iron oxide phases of magnetite, hematite, maghemite, or goethite.
19. The method of claim 17 wherein no traces of iron, iron oxides, or tin metal are observable in said nanoparticles by X-ray diffraction studies utilizing the Debye-Scherrer technique.
20. The method of claim 16, wherein said material is a film.
21. The method of claim 20, wherein said film is a thin film of Sn0.95Fe0.05O2.
22. An apparatus for detecting a gas comprising a gas inlet, a flow controller, a device which is configured to measure magnetic properties, and a ferromagnetic material, wherein the ferromagnetic material comprises Fe-doped SnO2 nanoparticles that exhibit room-temperature ferromagnetism.
23. The apparatus of claim 22 wherein the ferromagnetic material is Sn0.95Fe0.05O2.
US12/552,276 2004-07-30 2009-09-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism Abandoned US20100064771A1 (en)

Priority Applications (1)

Application Number Priority Date Filing Date Title
US12/552,276 US20100064771A1 (en) 2004-07-30 2009-09-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Applications Claiming Priority (4)

Application Number Priority Date Filing Date Title
US59820304P 2004-07-30 2004-07-30
US61270804P 2004-09-23 2004-09-23
US11/195,573 US7582222B2 (en) 2004-07-30 2005-08-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism
US12/552,276 US20100064771A1 (en) 2004-07-30 2009-09-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Related Parent Applications (1)

Application Number Title Priority Date Filing Date
US11/195,573 Continuation US7582222B2 (en) 2004-07-30 2005-08-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Publications (1)

Publication Number Publication Date
US20100064771A1 true US20100064771A1 (en) 2010-03-18

Family

ID=35787468

Family Applications (3)

Application Number Title Priority Date Filing Date
US11/195,571 Abandoned US20060060776A1 (en) 2004-07-30 2005-08-01 Method of sensing a gas by detecting change in magnetic properties
US11/195,573 Expired - Fee Related US7582222B2 (en) 2004-07-30 2005-08-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism
US12/552,276 Abandoned US20100064771A1 (en) 2004-07-30 2009-09-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Family Applications Before (2)

Application Number Title Priority Date Filing Date
US11/195,571 Abandoned US20060060776A1 (en) 2004-07-30 2005-08-01 Method of sensing a gas by detecting change in magnetic properties
US11/195,573 Expired - Fee Related US7582222B2 (en) 2004-07-30 2005-08-01 Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Country Status (2)

Country Link
US (3) US20060060776A1 (en)
WO (1) WO2006015321A2 (en)

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US11745169B1 (en) * 2019-05-17 2023-09-05 Unm Rainforest Innovations Single atom metal doped ceria for CO oxidation and HC hydrogenation/oxidation

Families Citing this family (20)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
WO2006015321A2 (en) * 2004-07-30 2006-02-09 Univ Boise State Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism and method of sensing a gas by detecting change in magnetic properties
US20080012004A1 (en) * 2006-03-17 2008-01-17 Mears Technologies, Inc. Spintronic devices with constrained spintronic dopant
US7625767B2 (en) 2006-03-17 2009-12-01 Mears Technologies, Inc. Methods of making spintronic devices with constrained spintronic dopant
US7471449B2 (en) * 2006-07-03 2008-12-30 Terahertz Technologies Llc Method and apparatus for generating Terahertz radiation with magnon gain medium and magnon mirror
US7430074B2 (en) * 2006-07-03 2008-09-30 Terahertz Technologies, Llc Generation of Terahertz waves
US7706056B2 (en) * 2006-07-03 2010-04-27 Terahertz Technologies Llc Modulation of terahertz radiation
US7986454B1 (en) 2006-07-03 2011-07-26 Terahertz Technologies Llc Tunable terahertz generator using a magnon gain medium with an antenna
US8031397B1 (en) 2006-07-03 2011-10-04 Terahertz Technologies, Llc Three-level magnon laser at room temperatures
US7836752B2 (en) * 2006-12-22 2010-11-23 Boise State University Magnetic gas sensor and methods using antiferromagnetic hematite nanoparticles
US7991480B2 (en) * 2007-08-28 2011-08-02 Cardiac Pacemakers, Inc. Medical device electrodes having cells disposed on nanostructures
US7894914B2 (en) * 2007-08-28 2011-02-22 Cardiac Pacemakers, Inc. Medical device electrodes including nanostructures
US8773118B2 (en) * 2008-07-11 2014-07-08 University Of Cape Town Magnetometer
US8427740B1 (en) 2010-03-10 2013-04-23 Terahertz Technologies Llc Modulation of terahertz radiation at room temperatures
TWI479667B (en) * 2010-09-13 2015-04-01 Nat Univ Tsing Hua Solar spectrum full band absorption of solar cells
US20120152335A1 (en) * 2010-12-20 2012-06-21 Shiu Hui-Ying Full-spectrum absorption solar cell
US8638440B1 (en) * 2012-06-27 2014-01-28 U.S. Department Of Energy Plasmonic transparent conducting metal oxide nanoparticles and films for optical sensing applications
US9097677B1 (en) 2014-06-19 2015-08-04 University Of South Florida Magnetic gas sensors
CN110455873B (en) * 2019-08-02 2022-02-18 湘潭大学 Method for improving performance of MoS2 gas sensor by adopting W doping
CN111285409A (en) * 2020-02-20 2020-06-16 复旦大学 Gas-sensitive nanomaterial based on single-layer ordered tin oxide nanometer bowl branched iron oxide nanorod structure, preparation process and application thereof
WO2022189504A1 (en) 2021-03-09 2022-09-15 Danmarks Tekniske Universitet Vibrating sample magnetometer (vsm) with a sample holder

Citations (27)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US3999947A (en) * 1974-10-11 1976-12-28 Matsushita Electric Industrial Co., Ltd. Reducing gas sensor and a method of producing the same
US4045178A (en) * 1975-11-08 1977-08-30 Matsushita Electric Industrial Co., Ltd. Reducing gas sensor
US4457161A (en) * 1980-10-09 1984-07-03 Hitachi, Ltd. Gas detection device and method for detecting gas
US4458242A (en) * 1980-10-25 1984-07-03 Matsushita Electric Works, Ltd. Gas detector
US4703646A (en) * 1985-05-30 1987-11-03 Siemens Aktiengesellschaft Operating method and sensor for gas analysis
US4731226A (en) * 1985-06-24 1988-03-15 Figaro Engineering Inc. Gas sensor
US4911892A (en) * 1987-02-24 1990-03-27 American Intell-Sensors Corporation Apparatus for simultaneous detection of target gases
US4935383A (en) * 1988-09-23 1990-06-19 The United States Of America As Represented By The Administrator Of The National Aeronautics And Space Administration Preparation of dilute magnetic semiconductor films by metalorganic chemical vapor deposition
US4957615A (en) * 1986-02-04 1990-09-18 Terumo Kabushiki Kaisha Oxygen sensor
US5296048A (en) * 1989-05-31 1994-03-22 International Business Machines Corporation Class of magnetic materials for solid state devices
US5372785A (en) * 1993-09-01 1994-12-13 International Business Machines Corporation Solid-state multi-stage gas detector
US5447054A (en) * 1990-03-02 1995-09-05 Eniricerche S.P.A. Gas sensors formed of thin tin oxide films, for gaseous hydro-carbon determination
US5756207A (en) * 1986-03-24 1998-05-26 Ensci Inc. Transition metal oxide coated substrates
US5788887A (en) * 1996-11-01 1998-08-04 E. I. Du Pont De Nemours And Company Antimony doped tin oxide electroconductive powder
US5788913A (en) * 1996-11-01 1998-08-04 E. I. Du Pont De Nemours And Company Processes to prepare antimony doped tin oxide electroconductive powders
US6132524A (en) * 1997-01-16 2000-10-17 Agency Of Industrial Science & Technology, Ministry Of International Trade & Industry Semiconductor magneto-optical material
US6246227B1 (en) * 1997-12-24 2001-06-12 James Hobby Sensor for measuring the magnetic characteristics of a gas
US6270741B1 (en) * 1997-02-28 2001-08-07 Kawasaki Jukogyo Kabushiki Kaisha Mitsubishi Corporation Operation management method of iron carbide production process
US6481264B1 (en) * 1998-10-26 2002-11-19 Capteur Sensors And Analysers Limited Materials for solid-state gas sensors
US6484563B1 (en) * 2001-06-27 2002-11-26 Sensistor Technologies Ab Method at detection of presence of hydrogen gas and measurement of content of hydrogen gas
US20030153088A1 (en) * 1999-01-15 2003-08-14 Dimeo Frank Micro-machined thin film sensor arrays for the detection of H2, NH3, and sulfur containing gases, and method of making and using the same
US6610421B2 (en) * 2000-09-08 2003-08-26 National Institute Of Advanced Industrial Science And Technology Spin electronic material and fabrication method thereof
US6632402B2 (en) * 2001-01-24 2003-10-14 Ntc Technology Inc. Oxygen monitoring apparatus
US6668627B2 (en) * 2000-10-04 2003-12-30 Swiss Federal Institute Of Technology Zurich Sensor apparatus with magnetically deflected cantilever
US20050100930A1 (en) * 2003-11-12 2005-05-12 Wang Shan X. Magnetic nanoparticles, magnetic detector arrays, and methods for their use in detecting biological molecules
US6996478B2 (en) * 1999-06-17 2006-02-07 Smiths Detection Inc. Multiple sensing system and device
US20060060776A1 (en) * 2004-07-30 2006-03-23 Alex Punnoose Method of sensing a gas by detecting change in magnetic properties

Patent Citations (28)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US3999947A (en) * 1974-10-11 1976-12-28 Matsushita Electric Industrial Co., Ltd. Reducing gas sensor and a method of producing the same
US4045178A (en) * 1975-11-08 1977-08-30 Matsushita Electric Industrial Co., Ltd. Reducing gas sensor
US4457161A (en) * 1980-10-09 1984-07-03 Hitachi, Ltd. Gas detection device and method for detecting gas
US4458242A (en) * 1980-10-25 1984-07-03 Matsushita Electric Works, Ltd. Gas detector
US4703646A (en) * 1985-05-30 1987-11-03 Siemens Aktiengesellschaft Operating method and sensor for gas analysis
US4731226A (en) * 1985-06-24 1988-03-15 Figaro Engineering Inc. Gas sensor
US4957615A (en) * 1986-02-04 1990-09-18 Terumo Kabushiki Kaisha Oxygen sensor
US5756207A (en) * 1986-03-24 1998-05-26 Ensci Inc. Transition metal oxide coated substrates
US4911892A (en) * 1987-02-24 1990-03-27 American Intell-Sensors Corporation Apparatus for simultaneous detection of target gases
US4935383A (en) * 1988-09-23 1990-06-19 The United States Of America As Represented By The Administrator Of The National Aeronautics And Space Administration Preparation of dilute magnetic semiconductor films by metalorganic chemical vapor deposition
US5296048A (en) * 1989-05-31 1994-03-22 International Business Machines Corporation Class of magnetic materials for solid state devices
US5447054A (en) * 1990-03-02 1995-09-05 Eniricerche S.P.A. Gas sensors formed of thin tin oxide films, for gaseous hydro-carbon determination
US5372785A (en) * 1993-09-01 1994-12-13 International Business Machines Corporation Solid-state multi-stage gas detector
US5788913A (en) * 1996-11-01 1998-08-04 E. I. Du Pont De Nemours And Company Processes to prepare antimony doped tin oxide electroconductive powders
US5788887A (en) * 1996-11-01 1998-08-04 E. I. Du Pont De Nemours And Company Antimony doped tin oxide electroconductive powder
US6132524A (en) * 1997-01-16 2000-10-17 Agency Of Industrial Science & Technology, Ministry Of International Trade & Industry Semiconductor magneto-optical material
US6270741B1 (en) * 1997-02-28 2001-08-07 Kawasaki Jukogyo Kabushiki Kaisha Mitsubishi Corporation Operation management method of iron carbide production process
US6246227B1 (en) * 1997-12-24 2001-06-12 James Hobby Sensor for measuring the magnetic characteristics of a gas
US6481264B1 (en) * 1998-10-26 2002-11-19 Capteur Sensors And Analysers Limited Materials for solid-state gas sensors
US20030153088A1 (en) * 1999-01-15 2003-08-14 Dimeo Frank Micro-machined thin film sensor arrays for the detection of H2, NH3, and sulfur containing gases, and method of making and using the same
US6996478B2 (en) * 1999-06-17 2006-02-07 Smiths Detection Inc. Multiple sensing system and device
US6610421B2 (en) * 2000-09-08 2003-08-26 National Institute Of Advanced Industrial Science And Technology Spin electronic material and fabrication method thereof
US6668627B2 (en) * 2000-10-04 2003-12-30 Swiss Federal Institute Of Technology Zurich Sensor apparatus with magnetically deflected cantilever
US6632402B2 (en) * 2001-01-24 2003-10-14 Ntc Technology Inc. Oxygen monitoring apparatus
US6484563B1 (en) * 2001-06-27 2002-11-26 Sensistor Technologies Ab Method at detection of presence of hydrogen gas and measurement of content of hydrogen gas
US20050100930A1 (en) * 2003-11-12 2005-05-12 Wang Shan X. Magnetic nanoparticles, magnetic detector arrays, and methods for their use in detecting biological molecules
US20060060776A1 (en) * 2004-07-30 2006-03-23 Alex Punnoose Method of sensing a gas by detecting change in magnetic properties
US7582222B2 (en) * 2004-07-30 2009-09-01 Boise State University Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism

Non-Patent Citations (4)

* Cited by examiner, † Cited by third party
Title
Coey et al., "Ferromagnetism in Fe-Doped SnO2 Thin Films", Applied Physics Letters, 23 February 2004, V 84 (8), PP 1332-1334. *
Fitzgerald et al, "SnO2 doped with Mn, Fe or Co: Room temperature dilute magnetic semiconductors," J. Applied Physics, 2004, 95(11 ), PP 7390-7392. *
Galatsis et al, "p- and n-type Fe-doped SnO2 gas sensors fabricated by the mechanochemical processing technique," Sensors and Actuators 2003, V B93, PP 562-565. *
Seki, K. et al., "A New Environmental Monitor Sensor Utilizing Thick-Film Magnetic Semiconductor", IEEE Transactions on Magnetics, September 1990, V 26 (5), PP. 2035-2037. *

Cited By (1)

* Cited by examiner, † Cited by third party
Publication number Priority date Publication date Assignee Title
US11745169B1 (en) * 2019-05-17 2023-09-05 Unm Rainforest Innovations Single atom metal doped ceria for CO oxidation and HC hydrogenation/oxidation

Also Published As

Publication number Publication date
US20060060776A1 (en) 2006-03-23
US20060060815A1 (en) 2006-03-23
US7582222B2 (en) 2009-09-01
WO2006015321A2 (en) 2006-02-09
WO2006015321A9 (en) 2007-04-26

Similar Documents

Publication Publication Date Title
US7582222B2 (en) Transition metal-doped oxide semiconductor exhibiting room-temperature ferromagnetism
Punnoose et al. Development of high-temperature ferromagnetism in Sn O 2 and paramagnetism in SnO by Fe doping
Varshney et al. Substitutional effect on structural and magnetic properties of AxCo1− xFe2O4 (A= Zn, Mg and x= 0.0, 0.5) ferrites
Seifert et al. Synthesis and magnetic properties of La-substituted M-type Sr hexaferrites
Liu et al. Synthesis and magnetic characterization of novel CoFe2O4–BiFeO3 nanocomposites
Martínez-Boubeta et al. Critical radius for exchange bias in naturally oxidized Fe nanoparticles
Dippong et al. Investigation of structural and magnetic properties of NixZn1-xFe2O4/SiO2 (0≤ x≤ 1) spinel-based nanocomposites
Yogi et al. Magnetic and structural properties of pure and Cr-doped haematite: α-Fe 2− x Cr x O 3 (0≤ x≤ 1)
Pathak et al. Structural and magnetic phase evolution study on needle-shaped nanoparticles of magnesium ferrite
Kuru et al. Dielectric, magnetic and humidity properties of Mg–Zn–Cr ferrites
Abdel-Khalek et al. Synthesis, structural and magnetic properties of La 1− x Ca x FeO 3 prepared by the co-precipitation method
Goya et al. Ferrimagnetism and spin canting of Zn 57 Fe2O4 nanoparticles embedded in ZnO matrix
Guigue-Millot et al. Control of grain size and morphologies of nanograined ferrites by adaptation of the synthesis route: mechanosynthesis and soft chemistry
Vučinić-Vasić et al. Zn, Ni ferrite/NiO nanocomposite powder obtained from acetylacetonato complexes
Ounacer et al. Substitutional effect of Mg2+ on structural and magnetic properties of cobalt nanoferrite
Cristobal et al. Mechanosynthesis and magnetic properties of nanocrystalline LaFeO3 using different iron oxides
Gupta et al. Physico-chemical analysis of pure and Zn Doped Cd ferrites (Cd1− xZnxFe2O4) nanofabricated by Pechini sol–gel method
Khanvilkar et al. Effect of divalent/trivalent doping on structural, electrical and magnetic properties of spinel ferrite nanoparticles
Mahmood et al. Physical characterization, optical properties, and magnetic interactions of cadmium-doped zinc ferrite nanoparticles
Zhang et al. Synthesis of magnetic manganese ferrite
Li et al. Size dependent magnetic and magneto-optical properties of Ni0. 2Zn0. 8Fe2O4 nanoparticles
Edelman et al. Formation, characterization and magnetic properties of maghemite γ-Fe2O3 nanoparticles in borate glasses
Kamran et al. Magnetic and dielectric properties of NiCrxFe2− xO4 nanoparticles
Fujii et al. Superparamagnetic behaviour and induced ferrimagnetism of LaFeO3 nanoparticles prepared by a hot-soap technique
Sanchez-Andujar et al. Enhanced magnetoresistance in the Ruddlesden–Popper compound Sr3Fe1. 5Co0. 5O6. 67

Legal Events

Date Code Title Description
STCB Information on status: application discontinuation

Free format text: ABANDONED -- FAILURE TO RESPOND TO AN OFFICE ACTION