nmr of nano particles

Upload: sohail

Post on 27-Feb-2018

247 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/25/2019 NMR of Nano Particles

    1/19

    NMR Techniques for Noble Metal Nanoparticles

    Lauren E. Marbella and Jill E. Millstone*

    Department of Chemistry, University of Pittsburgh, 219 Parkman Avenue, Pittsburgh, Pennsylvania 15260, United States

    ABSTRACT: Solution phase noble metal nanoparticle growth reactions arecomprised of deceptively simple steps. Analytical methods with high chemical,spatial, and temporal resolution are crucial to understanding these reactions andsubsequent nanoparticle properties. However, approaches for the character-ization of solid inorganic materials and solution phase molecular species areoften disparate. One powerful technique to address this gap is nuclear magneticresonance (NMR) spectroscopy, which can facilitate routine, direct, molecular-scale analysis of nanoparticle formation and morphology in situ, in both thesolution and the solid phase. A growing body of work indicates that NMRanalyses should yield an exciting complement to the existing canon of routine

    nanoparticle characterization methods such as electron microscopy and opticalabsorption spectroscopy. Here, we discuss recent developments in theapplication of NMR techniques to the study of noble metal nanoparticlegrowth, surface chemistry, and physical properties. Specically, we describe theunique capabilities of NMR in resolving hardsoft matter interfaces with both high chemical and high spatial resolution.

    1. INTRODUCTION

    Nuclear magnetic resonance (NMR) spectroscopy is atransformational molecular characterization tool that requireslittle perturbation of the analyzed system, while providingexceptional detail about the chemical environments ofconstituent atomic nuclei. These features make NMR especiallywell-suited for in situ analysis of chemical structure, reactions,

    and even dynamics in some cases. With this versatility, it is notsurprising that NMR analysis has been applied toawide varietyof systems1 ranging from large biomolecules2 to lithiumbatteries,3 in addition to its daily analytical use in organicsynthesis laboratories. The chemical resolution possible usingNMR is particularly attractive for characterizing both theformation and nal architecture of noble metal nanoparticles(NMNPs). To understand why NMR is promising for thesestudies one must clarify both what one may want to determineabout NMNP systems as well as the unique capabilities ofNMR in metallic materials.

    Analytical targets in the study of NMNPs range fromtracking molecular precursors during NP formation to particlesurface reorganization during catalytic reaction and many

    aspects of particle architecture, electronic properties, andsurface chemistry in between. Nanoparticle systems ofteninvolve a hardsoft matter interface between the solid surfaceof the particle core and pendant ligands (species ranging frommonoatomic ions to large macromolecules). This interfaceincludes many parameters of interest including surface elementcomposition, ligand shell composition, and ligand shellarchitecture. However, each of these features is difficult toresolve using classic surface and materials characterizationstrategies such as electron microscopy or photoelectronspectroscopy techniques.

    In this Perspective, we highlight reports, including examplesfrom our laboratory, in which NMR has provided crucialinsights into NMNP formation, morphology, and physicalproperties. First, we briey outline key NMR concepts in thestudy of NMNPs including NMR phenomena such as theKnight shift and Korringa behavior. In Section 2, we discussNMR analyses of NMNP formation and growth, and give

    examples of studies that monitor either resonances of the ligand(Section 2.1) or NMR-active metal nuclei within the metalprecursors (Section2.2). In the remainder of the Perspective,we discuss nal nanoparticle characterization both of theparticle itself (Section 3) as well as its resulting physicalproperties (Section4).

    In each section, we focus on NMR techniques that aregenerally accessible to the synthetic nanochemistry community.We also consider instrumental and physical limitations of NMRfor studying NMNPs where appropriate. At the end of eachsection, we include results obtained using more advanced NMRequipment and techniques. While these studies may requiremore expertise to execute, the results obtained are of broadinterest and therefore we highlight the method and the results.

    1.1. Basic Concepts for Metal Nanoparticle NMRSpectroscopy.While a detailed treatment of NMR theory isoutside the scope of this Perspective, the reader is referred toseveral excellent resources on the long and fruitful study ofmetals andmetal adsorbates using NMR and in particular thoseby Slichter,4 van der Klink and Brom,5 and Oldeld andWieckowski.6 Here, we outline selected NMR concepts. Theconcepts are chosen to be useful for interpreting the results

    Received: December 31, 2014Revised: February 19, 2015

    This Perspective is part of the Up-and-Comingseries.

    Perspective

    pubs.acs.org/cm

    XXXX American Chemical Society A DOI: 10.1021/cm504809c

    Chem. Mater. XXXX, XXX, XXXXXX

    This is an open access article published under an ACS AuthorChoice License, which permitscopying and redistribution of the article or any adaptations for non-commercial purposes.

    http://localhost/var/www/apps/conversion/tmp/scratch_7/pubs.acs.org/cmhttp://dx.doi.org/10.1021/cm504809chttp://pubs.acs.org/page/policy/authorchoice_termsofuse.htmlhttp://pubs.acs.org/page/policy/authorchoice_termsofuse.htmlhttp://pubs.acs.org/page/policy/editorchoice/index.htmlhttp://dx.doi.org/10.1021/cm504809chttp://localhost/var/www/apps/conversion/tmp/scratch_7/pubs.acs.org/cm
  • 7/25/2019 NMR of Nano Particles

    2/19

    summarized in this Perspective as well as for appreciating thescope of possible contributions using NMR spectroscopy inNMNP systems. In particular, we focus on phenomena uniqueto NMR of metals, adsorbates on metals, and nuclei at a hardsoft matter interface.

    Consider an NMR spectrum of 195Pt nuclei in Pt nano-particles or a 1H NMR spectrum of H2molecules adsorbed to

    their surface. In both cases, conduction electrons in the Ptmetal will have a dramatic inuence on the resulting NMRspectra because of nuclear coupling to conduction electrons atthe particle surface (dened as hyperne coupling). In bulkmetals, it is well-known that hyperne coupling of nuclear spinsto the unpaired conduction electrons in the metal results in adramatic change in NMR frequency termed the Knight shift, K.Observation of the Knight shift was rst reported in 1949 whenW. D. Knight noticed that the NMR signal for Li, Na, Al, Cu,and Ga metals resonated at a different frequency from that ofthe same element in a nonmetallic environment (chloridesalts).7

    Like chemical shift, the Knight shift is sensitive to the localelectronic environment surrounding the nucleus, and the

    magnitude of the shift is a sensitive probe of the electronicstructure of metals, semiconductors, and superconductors.8

    Therefore, exciting experiments such as probing the changefrom molecular to metallic electronic structure in NMNPs ispossible using NMR techniques. Further, metallic propertiescan be studied by evaluating the Korringa behavior of materialsthrough NMR measurements. In the Korringa relation,9 K isrelated to the temperature,T, at which the NMR measurementis performed and the longitudinal relaxation time constant,T1,of the material where K2TT1is a constant. As a result, T1andT1 display a linear relationship with one another.10 NMRnuclei that exhibit this temperature correlation with T1suggestmetallic electronic structure within the material analyzed. Boththe slope of the temperature dependent relaxation as well as the

    magnitude of the Knight shift can give information about theevolution of particle electronic structure as a function of particlesize, shape, composition, and surface chemistry.8

    Surface chemistry is particularly interesting because NMNPsprepared by wet chemical techniques are often capped withorganic molecules. By combining the unique impact of metalconduction electrons on ligand nuclei with traditional NMRspectroscopy techniques, NMR investigation of these cappingligands can provide detailed insight into properties of theparticle core (e.g., electronic structure, atomic composition, orcompositional architecture) as well as important aspects of itsligand shell including ligand identity, arrangement, anddynamics.

    For each nucleus observed, the measured resonance will be

    inuenced by both macroscopic

    and microscopic

    forcesacting on it.11 At the macroscopic level, the homogeneity andstrength of the applied magnetic eld, molecular tumbling, andmagnetic susceptibility of the material can each inuenceobserved spectral features. Many NMR techniques have beendeveloped to either mitigate the impact of these macroscopiceffects or to leverage them as in solid state NMR (ssNMR).Microscopicfactors include the electronic environment of thenuclei as well as the inuence of neighboring nuclear andelectronic spin interactions. These microscopic interactionsmay be deduced from several spin-1/2 NMR gures of merit,including chemical shift, line width, relaxation times, andchemical shift anisotropy (CSA).

    For a given chemical environment, several of thesespectroscopic parameters may be inuenced simultaneously.Further, because NPs incorporate components from bothmolecular chemistry (e.g., organic capping ligands) and solidphase materials (e.g., metallic core), methods may be used orcombined from both traditional solution phase NMR character-ization as well as more advanced ssNMR techniques. In orderto ascertain how different spectral features arise in spectraobtained from these techniques, we begin by discussing atraditional solution phase spin-1/2 1D NMR spectrum andfocus on a single nuclear site,j. In a typical spectrum of a smallmolecule, sharp resonance lines of chemical shift, iso,j, areobserved at a frequency (in ppm) that is proportional to theexternal magnetic eld, B0.

    = +B (1 )j j j0, 0 iso, (1)

    In eq1,0,j is the chemically shifted Larmor frequency for sitej, and j is the gyromagnetic ratio of nucleus j. Here, thedirection of the external magnetic eld is in thez-direction of athree-coordinate axis. Each nuclear site may exhibit a uniquechemical shift as the result of differences in chemical

    shielding.12

    The chemical shielding interaction is composedof both a diamagnetic and a paramagnetic contribution thatresults in frequency shifts.13 The magnitude of the chemicalshielding depends on the molecular electronic structure and theorientation of the molecule with respect to B0.

    14 In solution,rapid reorientation of the small molecule with respect to B0allows it to sample all orientations on a time scale that averagesthe CSA and dipoledipole coupling to zero. This averagingresults in anisotropicchemical shift and sharp resonance lines inthe solution phase NMR experiment (Figure1A).

    However, a static 13C ssNMR spectrum of the same smallmolecule shows a broad powder pattern. This apparent loss inspectral features is due to spin interactions such as CSA anddipoledipole coupling which are not averaged out in the solid

    state because of restricted molecular motion. In theory, theeffects of both CSA and dipoledipole coupling can be

    Figure 1.1D 13C NMR spectra of 11-mercaptoundecanoic acid in (A)D2O, (B) the solid state with MAS = 5 kHz and

    1H decoupling = 80kHz, (C) D2O appended to Au NPs (d= 2.2 nm), and (D) the sameNPs in the solid state. Asterisks in (B) denote spinning sidebands frommoderate MAS speeds = 5 kHz.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    B

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    3/19

    removed from ssNMR spectra via magic-angle spinning (MAS)in combination with high power decoupling (Figure1B).

    When small molecules are attached to a metal nanoparticle,they are at the interface between solution and a solid support.Therefore, if the same small molecule depicted in Figure 1A isappended to a AuNP and the same 1D 13C NMR spectrum isrecorded, several spectral differences will be apparent: linebroadening of the resonances is observed, the chemical shiftmay be altered, and some resonances may disappear completely(Figure1C). These changes become more dramatic when theparticles are dried to a solid state (Figure 1D). For spin-1/2nuclei, such as the 13C solution phase spectra in Figure 1C, thespin interactions that arise from being in a more solid-likeenvironment can be briey explained as a reintroduction ofanisotropic spin interactions. For example, CSA can arise fromthe different frequencies associated with each differentorientation of the same molecule with respect to B0. If themolecule does not reorient at a rate greater than the absolutemagnitude of the CSA (which can be the case when attached toa nanoparticle substrate because of restricted molecularmotion), inhomogeneous line broadening is observed in theNMR spectrum.

    Likewise, dipoledipole coupling can also be a source of linebroadening and is the through-space interaction between theinduced magnetic moments of neighboring spins. Measuringdipoledipole coupling constants can provide detailedinformation about the structure and arrangement of ligandsattached to NMNPs in both the solid and solution phase. Forexample, the effect of dipoledipole coupling on nuclear spinrelaxation results in nuclear Overhauser effects that can bemeasured by NMR and provide information on the distancesbetween nuclei (termed nuclear Overhauser effect spectroscopy(NOESY)). Using specialized pulse sequences, spin inter-actions can be selectively reintroduced to learn about themolecular environment of nanoparticle capping ligands,including the degree of crystallinity, orientation, and supra-

    molecular architecture.Depending upon the size of the underlying nanoparticle,

    molecular tumbling of pendant ligands may be greatly reduced.The attenuation of molecular tumbling rates via attachment torelatively large particle surfaces results in a reintroduction ofCSA and dipoledipole coupling spin interactions as discussedabove. Changes in the magnitude of spin interactions (e.g., byxing molecules in close proximity or near a metal center) canalso inuence the spinlattice relaxation (or longitudinalrelaxation, T1) and spinspin relaxation (or transverserelaxation, T2). Here, T1 is dened as the time constant forspins to reach thermal equilibrium in the presence of B0.Similarly,T2is the time constant that describes the dephasingof spin polarization associated with single quantum coherences

    in the transverse plane. In particular, T2 is related to theobserved NMR full width at half-maximum (fwhm) by thefollowing equation:

    =

    Tfwhm

    1

    2 (2)

    Another spin interaction that will be mentioned is thequadrupolar interaction. Quadrupolar nuclei are nuclei withspin > 1/2 and can have either integer or half-integer spins. Inthese cases, the nucleus has an electric quadrupole momentwhich is coupled to the surrounding electric eld gradient,producing the quadrupolar interaction. Quadrupolar interac-tions are commonly characterized by nuclear quadrupolar

    coupling constants. It is important to note that, in manysystems, especially solids where molecular motion is restricted,the magnitude of quadrupolar coupling can surpass themagnitude of the dipoledipole coupling interactions, produc-ing NMR lineshapes that are dominated by quadrupolarinteractions.15 Here, we will focus mainly on the quadrupolarinteraction as it applies to deuterium NMR spectroscopy, wheredeuterium is spin-1 and has a relatively small quadrupolemoment.

    2. NUCLEAR MAGNETIC RESONANCECHARACTERIZATION OF METAL NANOPARTICLEFORMATION AND GROWTH

    During the formation of NMNPs, both chemical change andphase transformations occur. Following the evolution ofmolecular precursors into a nal NP solid phase requiresmethods that can capture chemical and physical trans-formations in real time with molecular resolution. NMRapproaches to study growth have employed a broad range oftechniques, as well as combined NMR with other analyticalmethods such as thermogravimetric analysis (TGA), Ramanspectroscopy, and differential scanning calorimetry (DSC). Tostudy NMNP growth, both the nuclei of NP ligands as well asthe metal nuclei themselves can be monitored.

    2.1. NMR Observation of Nanoparticle LigandResonances during Synthesis.First, we consider the useof NMR to monitor the chemical environment, reaction rates,and dynamics of ligands used in NMNP syntheses. Onenanochemical reaction mechanism that has beneted signi-cantly from NMR analysis is the two-phase BrustSchiffrinsynthesis.16 Briey, this synthesis involves the transfer ofaqueous HAuCl4into an organic solvent such as toluene via aphase transfer agent such as tetraoctylammonium bromide(TOAB), and a thiolated ligand (RSH) solubilized in theorganic phase. A reducing agent such as NaBH4(aq) isintroduced and thiolate-terminated AuNPs are produced(diameter d 15 nm).16 In general, the mechanism offormation is thought to begin by an initial reduction of Au(III)species to Au(I) via oxidation of RSH ligands. The resultingspecies are then fully reduced by NaBH4and particle nucleationis induced. On the basis of dynamic light scattering (DLS)measurements and the sensitivity of the nal product size tometal:RSH ratio, it has been thought that metalsulfur bondformation occurred after RSH addition and that polymericAu(I)-thiolate species were the active precursor to thiolate-capped NPs produced via the two-phase approach.

    Using 1D 1H NMR analyses of precursors at various Au:RSHratios, Lennox and co-workers found that, after thiol addition,Au(III) is indeed reduced to Au(I) and that RSH is oxidized tothe disulde but that there was no evidence of metalsulfur

    bond formation prior to the introduction of NaBH4 (Figure2A,B).17 Instead, Lennox postulates the formation of TOA+[AuX2]

    species and supports this assignment via chemical shiftdifferences in the TOA+ resonances after the addition ofHAuCl4 and its transfer to the organic phase. TOA

    + protonresonances (protons on the carbons adjacent to the nitrogen)were consistent with fast anion exchange between coordinationof TOA+ to [AuX2]

    and Br. Similar metalsurfactantprecursors were subsequently identied for analogous Ag andCu two-phase NP syntheses.17

    This work led to a proposed inverse micelle mechanismbyTong and co-workers18 where the metal ion coordinationcomplex is sequestered in an inverse micelle of TOAX (X = Br

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    C

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    4/19

    or Cl). Both the size of the micelle and the chemicalenvironment inside play crucial roles in the ultimate size andstability of AuNPs formed, even in the absence of an RSHligand. Tong et al. support this mechanism via both NMR andsynthesis experiments. Using 1D 1H NMR, the authors followthe water resonance and show chemical shift valuesconsistentwith sequestration of water inside a TOAX micelle19 or othersupramolecular architecture. Further, the TOA+[AuX2]

    precursor could be prepared in high purity according to aliterature procedure, eliminating the need for RSH reduction ofTOA+AuX4]

    . Then, the thiolated capping ligand could beadded post-NaBH4 addition and indistinguishable particlescould be obtained for Au, Ag, or Cu cores (Figure1C).

    This particle formationpathway differs from observations ina one-phase synthesis,20,21 involving the same reagents with theexception of the phase transfer agent TOAB, which is no longerneeded and typically omitted. In the one-phase synthesis,metalsulfur bonds are observed before NaBH4 addition,which is consistent with results from our own laboratory. Cliffeland co-workers studied the one-phase aqueous synthesis oftiopronin-capped AuNPs and used a combination of 1H NMR,TGA, matrix assisted laser desorption ionization massspectrometry (MALDI-MS), and pair distribution function(PDF) analysis to identify precursor species.22 The authorsshowed that a possible precursor to AuNPs in the tioproninsynthesis was a Au(I)thiolate tetramer with distinct optical

    signatures, and these observations are consistent withtheoretical predictions of pre-nucleationspecies.23,24 Interest-ingly, features of the Au(I)thiolate tetramer are also presentin the nal colloid (d = 23 nm), suggesting that theintermediates may also be present as capping moieties on theparticle surface. Studies of analogous syntheses usingphosphine-based ligands have been studied using 31P ssNMRspectroscopy and indicate that Au(I) ligand complexes alsoform prior to reduction and subsequent particle formation.25

    Taken together, these NMR studies suggest that there arefundamental differences between the formation pathways ofsmall NMNPs (d = 15 nm) synthesized by one-phase andtwo-phase methods, despite the seemingly similar protocols.These distinctions are important contributions to establishingwell-controlled, easily tailored, and high yielding NMNPsyntheses.

    Another canonical NMNP synthesis studied by solutionphase 1H NMR is the citrate reduction of Au(III) at elevatedtemperature (100 C), commonly referred to as theTurkevich26 or Frens27 method. In this synthesis, citrate actsas both a reducing agent for Au(III) as well as a capping ligandfor the nal NPs. By altering either the solution pH (andpresumably the speciation of both the Au precursor and/or thereduction potential of the citrate) or the ligand to metal ratio,the nal particle size and size distribution can be modied (d10 to 100 nm). In order to correlate variations in precursorchemistry with particle products, Bruylants and co-workersmonitored the Turkevich synthesis at constant citrate:Au(III)ratio (5:1) at various pH values (pH = 3, 4.5, 7, 9, and 12)using a combination of UVvisible absorption and NMRspectroscopy as a function of time.28

    In this work, the authors identied that the narrowest particlesize distribution was obtained from reactions performed at pH7. At pH 7, a majority of the citrate ligand was found by 1HNMR to be deprotonated (pKa = 3.1, 4.8, 6.4) and the Auspecies was present as [AuCl2(OH)2]

    as determined by UV

    visible spectroscopy. Under these conditions, dicarboxyacetoneappeared in the 1H NMR spectrum within 1 min of reactioninitiation, where dicarboxyacetone is the primary oxidationproduct of citrate and itself a reducing agent. However, whenthe reduction of Au cations is performed with dicarboxyacetonealone, particle size distribution increases, indicating that thereaction is sensitive to kinetics and/or progresses by a differentroute than when using citrate.

    In addition to the appearance of the dicarboxyacetone peak, anew, broad peak at the base of the free citrate peaks wasobserved. 2D 1H1H correlation spectroscopy (COSY)analysis indicated that the peaks were consistent with citratein fast exchange with two different chemical environments onthe NMR time scale. When the diffusion coefficients of the

    various citrate peaks were measured by diffusion orderedspectroscopy (DOSY), citrate aggregates containing Au atoms(either Au(I) or Au(0); not yet known) were observed.28

    These Aucitrate aggregates are consistent with the supra-molecular metalligand assemblies proposed to explainnucleation and growth of AuNPs using the Turkevichapproach.26

    NMR techniques may also be used to follow NMNPsyntheses that use a gas reduction approach. In one example,Chaudret et al. characterize AuNP formation (d= 4.7 0.9nm) using a combination of solution phase and ssNMRtechniques to monitor the CO(g)-reduction of a gold precursorin the presence of amine-functionalized capping ligands.29

    Figure 2.(A) 1H NMR spectra of TOAX + TOA+[AuX4] titrated

    with increasing equivalents of dodecanethiol from top to middle aswell as spectra of pure TOAB, dodecanethiol, and dodecyl disuldefree in solution. (B) Diagram describing possible nanoparticleprecursors in the rst two steps of a two-phase BrustSchiffrinsynthesis based on spectra in (A). (C) Proposed scheme fornanoparticle formation based on similar NMR analyses to (A) andRaman spectroscopy. (A) and (B) adapted with permission from ref17. Copyright 2010 American Chemical Society. (C) adapted withpermission from ref18. Copyright 2011 American Chemical Society.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    D

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    5/19

    Solution phase 1H NMR measurements showed that thereaction shared mechanistic features with the one-phase BrustSchiffrin synthesis. For example, after combining hexadecyl-amine (HDA) with Au(I)Cl(tetrahydrothiophene) (THT), aAu(I)ClHDA coordination complex is formed. Resultingspectral features support this assignment including thedisappearance of the triplet from the protons on the carbonadjacent to the amine group, coincident appearance of newresonances that correspond to Au(I)ClHDA, and theappearance of resonances that correspond to free THT. Afterreduction of the Au(I)ClHDA complex by CO(g), AuNPs areformed and were characterized using ssNMR MAS techniques(to mitigate spectral line-broadening due to the increasing sizeof the NP core). 13C ssNMR spectroscopy indicated that, afterreduction, 1,3-dihexadecylurea was formed via carbonylation ofHDA. As a result, the nal particle product was terminated witha binary ligand shell composed of both amine and carbamideligands.

    NMR has also been used to investigate how early NP nucleimay be stabilized but continue to grow. Studying an IrNPsystem, Finke and co-workers used a combination of1H and 2HNMR spectroscopy to understand how ionic liquid media areable to stabilize transition metal NPs by monitoring the ionicliquid chemistry during synthesis.30 The authors found thatIr(0) NPs (d= 2.1 nm) reacted with imidazolium-based ionicliquids in order to form surface-bound carbenes, suggesting thatchemical reaction of the solvent can create molecular stabilizersfor the growing particle surface. Solvent and other reactionbyproducts have been implicated as stabilizingagents in otherNP syntheses as well. For example, Polte et al.31 have shownthat borate byproducts may stabilize Au and Ag NPs on thetime scale of hours during and after their formation when usingNaBH4as a reductant in the absence of other ligand reagents.

    In addition to determining the chemical evolution of ligandprecursors and their role in particle growth, NMR has also beenused to probe the role of capping ligands in the emergence of

    particle shape. For example, Gordon and co-workers used1H13C cross-polarization (CP)-MAS NMR spectroscopy todetermine the surface composition of poly(vinylpyrrolidone)(PVP, MW = 29 kDa)-capped AgPt nanocubesand octahedra(d = 68 nm) before catalytic evaluation.32 As may beexpected, when more Ag precursor is added to the synthesis,more Ag is incorporated into the nal particle architecture. Foreach new composition, dramatically different 1H13C CP-MASNMR spectra are observed, each of which deviates from thespectrum of pure PVP. Upon increasing Ag addition to the nalparticle, 13C resonances that correspond to hydrocarbons andamines begin to appear. At the same time, characteristic 13Cresonances of PVP, specically those from the 5-memberedring in the polymer repeat unit, disappear. Combined, these

    results suggest that Ag promotes bond cleavage to formhydrocarbon and amine fragments from the original PVPligandan interesting observation of changes in ligandchemistry initiated by the forming particle itself. Importantly,these spectroscopic changes in ligand chemistry as a function ofAg addition correlated not only with alloy composition but alsowith overall alloy shape, suggesting that metalPVP reactionsmay bias surface growth rates and inuence the nal particlemorphology.

    2.2. NMR Observation of Metal Nuclei during Syn-thesis. In addition to observing ligand chemistry during thesteps of NMNP synthesis, NMR-active metal nuclei allow directobservation of metal precursor behavior as well. Unfortunately,

    while 100% naturally abundant, the large quadrupole momentof the 197Au nucleus precludes direct observation via NMR withcurrent methods, although coupling to heteroatoms, such as31P, has been reported.33 Luckily, other metal nuclei have morefavorable NMR properties, such as 195Pt and 109/107Ag. WhileAg NMR holds promise for understanding silver NP precursorchemistry, both isotopes suffer from low sensitivity because of

    low gyromagnetic ratios and therefore generally requirespecialized low-gamma probe hardware for NMR observation.Further, Ag nuclei typically exhibit extremely long T1values (onthe order of hours) making data acquisition time-consuming.

    Despite these difficulties, understanding the magneticresonance properties of inorganic silver compounds, many ofwhich are NP precursors, is an area of active research34 and theapplications to NP synthesis are beginning to be explored. Forexample, Liu, Saillard, and co-workers have used a combinationof NMR spectroscopies (1H, 2H, 31P, 77Se, and 109Ag), UVvis,ESI-MS, FTIR, TGA, elemental analysis, single crystal X-raydiffraction, and density function theory (DFT)) to monitor theconversion of Ag(I) salts into hydride-centered Ag clusters(numberofatoms ranging from 7 to 10) and larger Ag NPs (d= 30 nm).35 Here, 109Ag NMR was used to locate the hydrogenatom within the Ag clusters. Figures of merit such as thecoupling constant between Ag and H were consistent whenmeasured either by 1H NMR or 109Ag NMR (JHAg= 39.4 Hzand JAgH 39.7 Hz, respectively) indicating a robust structureassignment.

    To date, the majority of work studying NMNP formation viaNMR of metal nuclei has focused on Pt and specically on Ptprecursors. One of the rst studies was reported by Murphyand co-workers and used a combination of solution phase 195Pt,1H, and 13C NMR to determine the dynamics of ligandexchange between [PtCl4]

    2 and poly(amidoamine)(PAMAM) dendrimers during dendrimer-mediated PtNPsynthesis.36 195Pt NMR, in particular, was used to identify thespeciation of the metal precursor upon binding to thedendrimer as well as track Pt precursor uptake into thetemplate branches. The goal of the study was to understand thechemical environment of the Pt(II) precursor (starting material= K2PtCl4) before reduction to Pt(0) by NaBH4 in thepresence of either generation 2 (G2-OH) or generation 4 (G4-OH) PAMAM dendrimers. In both G2-OH and G4-OH cases,the 195Pt NMR signal intensity assigned to [PtCl4]

    2 decreased,while new peaks arose as Cl ligands were replaced by nitrogen-containing sites from within the dendrimer. On the basis ofchemical shift analyses, the number and type of nitrogensubstitution could be deduced, with each successive aminesubstitution leading to a chemical shift change of 261 ppmand each successive amide substitution leading to a chemicalshift change of333 ppm.

    In addition to chemical shift distribution, the total Pt(II)signal could also provide information about the sample. Forexample, in the case of G2-OH, 50% of the total 195Pt NMRsignal disappeared during uptake, and this decrease wasattributed to a black Pt(0) precipitate that formed during thecourse of uptake (2 days). In the G4-OH system noprecipitate is observed over 10 days, but the 195Pt NMR signalintensity is attenuated by80%, indicating that the discrepancymay be due to signal dephasing of Pt nuclei in areas ofrestricted molecular motion such as inside the dendrimer orfrom aggregation of more than one dendrimer. Overall, theauthors found that Pt(II) uptake, speciation, and reactivity wasa complex result of the dendrimer architectures, correlating

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    E

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    6/19

    with both the impact of ligand exchange about the Pt(II) centeras well the sterics of the dendrimer as a whole.

    Recently, we have studied the role of Pt(IV) speciation onreaction pathways in Pt-containing, mixed metal NP syntheses.Specically, we examined the inuence of Pt(IV) speciation onthe deposition of Pt metal onto colloidal AuNP substrates(Figure3).37 Here, Pt(IV) speciation was controlled by addingNaOH to the H2PtCl6 precursor solution which we used as astrategy to control the extent of Pt hydrolysis that occurs inaqueous solutions of H2PtCl6. We then monitored the resulting

    ligand substitution of the Pt center as a function of pH viasolution phase 195Pt NMR spectroscopy.Monitoring ligand substitution of these complexes via 195Pt

    NMR is particularly attractive because chemical shift assign-ment can be supported by isotopologue analysis of the pendantchloride ligands. In this analysis, the isotopologue distributionof 35Cl/37Cl species in the [PtCl6]

    2 and [PtClxLy]n (where L

    = H2O or OH) can be extracted from deconvolution of NMR

    peaks (Figure 3A, insets). Specically, at high eld strengths(B0 = 14.1 T),

    37Cl substitution leads to an upeld shift of0.17 ppm.38 The relative populations of isotopologues in thespectra directly correlate with the natural abundance of therespective chlorine isotopes. Further, for monosubstituted[PtCl5L]

    n species, resolution of both 35/37Cl isotopologues

    and isotopomers is possible (isotopomers are represented bythe same color in the peak tting in Figure 3). This analysisprovides both a spectroscopic ngerprint for Pt(IV) complexesin solution as well as a relative quantication of various Ptspecies in solution.39

    Ultimately, Pt(IV) speciation was found to correlate wellwith nal NP morphologies, and the pH of the original Pt(IV)precursor solution was a reliable method to modulate thisspeciation parameter. Particularly interesting was the impact ofspeciation at high precursor solution pH. At pH 8.6, 67% of thePt(IV) complexes are [PtCl6]

    2 and 33% are the monosub-stituted [PtCl5(OH)]

    2. When Pt(IV) precursor solutions atthis pH are used for Pt deposition on the AuNP substrates, the

    majority of AuNPs are oxidized via a galvanic replacementmechanism and the synthesis results in the formation offramelike AuPt alloy nanostructures (Figure3D). Conversely,when Pt(IV) precursor solution pH is low, reduction of themetal cation occurs primarily via externally added reducingagent oxidation (in this case, ascorbic acid), as evidenced bylack of oxidation in the Au nanoprism substrate (Figure 3C).The difference in Pt(IV) reduction pathway is consistent withprevious electrochemical studies of [PtCl6]

    2 in water, whichnd that OH substituted halide complexes are more readily

    reduced.40

    Therefore, controlling Pt(IV) speciation (andthereby metal precursor reduction potential) is a synthetichandle with which to mediate whether Pt(IV) reduction willoccur via oxidation of the small molecule reducing agent(leading to deposition on top of the existing particle substrate)or through oxidation of existing NPs in solution (leading tocage-like, hollow, or frame nanostructures). These insightsprovide important mechanisms for tunability in the synthesis ofPt-containing NP products which are of interest in manyapplications including fuel cells41 and data storage.42

    Overall, the literature suggests that unprecedented, molec-ular-scale information can be gained by using NMR to monitorthe chemical conversion of NP precursors in both the solutionand the solid phase, under diverse reaction conditions, and for

    several diff

    erent metal identities. These insights should lead to adeeper understanding of NMNP synthesis mechanisms andultimately provide a robust foundation for future NP synthesisdesign.

    3. NUCLEAR MAGNETIC RESONANCECHARACATERIZATION OF NOBLE METALNANOPARTICLES

    In addition to providing critical insights into the formation ofNMNPs, NMR techniques may also be used to understandseveral aspects of the nal nanoparticle architecture. NMRstands out here, because it is able to resolve moleculararchitectures at the surface of the solid phase nanoparticle. In

    Figure 3. 195Pt NMR spectra of Pt(IV) precursor solutions at pH 1.8 (A) and 8.6 (B). Corresponding TEM images of particle motifs of Ptdeposition on Au nanoprism substrates when Pt(IV) precursor solution pH is 1.8 (C) and 8.6 (D). Insets in (A) show 195Pt NMR chemical shiftassignments for monosubstituted species [PtCl5(aq)]

    (left) and [PtCl6]2 (right) using isotopologue analysis at B0 = 14.1 T. In the case of

    monosubstitution, isotopomers could also be resolved upon peaktting. Black lines represent experimental spectra, colored lines represent the peakts for each isotopologue and/or isotopomer, and dashed gray lines represent the sum of the peak t. Adapted with permission from ref37.Copyright 2014 American Chemical Society.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    F

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    7/19

    this section, we discuss the use of NMR techniques to elucidatefeatures of both the nanoparticle core and the nanoparticleligand shell with high spatial and chemical resolution.

    3.1. Metal Nanoparticle Surface Chemistry: SmallMolecule Ligand Shells. 3.1.1. Ligand Identity andQuantity. An important aspect of ligand shell characterization

    is monitoring the process and

    nal products of ligand exchange.These studies require the ability to evaluate both ligand identityas well as quantity. NMR spectroscopy is particularly well-suited for these requirements, because of the chemicalresolution offered by chemical shift and the direct relationshipbetween NMR signal integration and spin population. Indeed,solution phase 1H NMR spectroscopy has been used toquantify the type and amountof ligands on quantum dots andmetal oxide nanoparticles,4346 and similar approaches havebeen used to assess the relative ratios of ligands in mixedmonolayer systems appended to NMNPs.4750 However, inorder to achieve quantitative results, NMNP cores typicallymust be digested due to line broadening effects that becomeincreasingly problematic at larger particle sizes (vide supra).51,52

    Perhaps more importantly, with all bulk ligand quanti

    cationstrategies to date (including those using NMR), the accuracy ofligand density values is fundamentally limited by the sizedistribution of the nanoparticle cores analyzed.

    Using NMR to monitor ligand exchange in real time wasperformed without particle digestion to gain information onligand exchange. At very small particle sizes (generally

  • 7/25/2019 NMR of Nano Particles

    8/19

    techniques can be used to probe these interactions. Forexample, 1D and 2D 1H high resolution (HR)MAS NMRtechniques showed the presence on stacking betweenaromatic ligands appended to a AuNP surface (orientation ofthe stacking with respect to the particle surface was notresolved).56 One may also systematically monitor ligandposition by using either the particle surface itself57 or site-specic paramagnetic lanthanide labels,58,59both of which act asa spectroscopic ruler by dephasing resonances closest to theunpaired electron (for more information about distance-dependent NMR signal dephasing using a variety of methods,the reader is referred to the work of Solomon60 andBloembergen61). Both strategies have been used to assessparameters such as proteinnanoparticle binding sites onspecic residues57 and molecular position-mapping of organiccapping ligands,58,59 respectively.

    Although limited distances (5 ) can be measured withsolution phase NMR techniques, detailed structural informationon various ligand shell architectures and arrangement can stillbe ascertained. Recently, Stellacci and co-workers presented a1H NMR method to determine ligand shell morphologies bypredicting 1H chemical shift trends as a function of ligand shellcomposition (also measured with 1H NMR, but with digestedparticles) in combination with NOE cross peak patterns(Figure 5).62 Using this method, the authors were able to

    determine the difference between randomly mixed, Janus, and

    patchy ligand shell morphologies on AuNPs (d = 25 nm)capped by binary mixtures of aromatic and aliphatic molecules.Further, 1H NMR line narrowing of specic ligand resonancesprovided spectroscopic evidence for structural defects asso-ciated with particular morphologies.

    To extract architectural information from this combination of1D 1H NMR spectra and NOE patterns, two importantexperimental conditions must be met. First, the binary ligandshell composition must be tunable across all composition space(i.e., from 0 to 100% of each ligand). This control is needed inorder to observe the changes in chemical shift associated withthe ligand mixtures as shown in Figure 5AC. Second, the 1HNMR resonances of the constituent ligands must be well-

    resolved to observe the cross-peaks shown in Figure5DF (i.e.,distinct and preferably very well-resolved from one another onthe chemical shift spectrum). The rst limitation can beovercome as a more robust understanding of ligand shellchemistry evolves, with a signicant amount of progressattributed to the information gained from NMR studies (videsupra). Additionally, the need to study mixed ligand systemswith well-resolved resonances can be mitigated by moving tohigher eld strengths and/or other observable nuclei with largerchemical shift ranges than traditional 1H NMR, such as 31P or13C.

    Additional spatial information on nanoparticle ligand shellscan be gained by using advanced ssNMR techniques, such asrotational echo double resonance (REDOR).63 Strong dipolarcouplings between neighboring nuclei lead to broad NMR linesin the solid state that can be averaged out with MAS. Due tolong-range order observed in solids, selective reintroduction ofdipolar couplings with REDOR allows for the measurement ofmuch longer distances between nuclei (25 ) than in thesolution phase, in which dipolar couplings are averaged out bymolecular tumbling.

    One example of this technique was used to distinguishbetween bilayer formation via a disulde linkage vs a hydrogenbonding interaction with important implications for cysteine-mediated protein binding to AuNP surfaces. Here, Gullion andco-workers used {1H}13C{15N} REDOR to selectivelyreintroduce heteronuclear dipolar couplings on uniformlylabeled 13C, 15N L-cysteine and L-cystine capped-AuNPs (d=6.6 nm).64 REDOR measurements along with 1H13C CP-MAS NMR analysis showed that the thiol-group of the L-cysteine molecule was chemisorbed to the Au surface andformed an initial ligand layer. Thiol anchoring of the initialmonolayer exposed charged amino and carboxyl groups on thezwitterionic L-cysteine, to which a second L-cysteine layercoordinated via hydrogen bonding. 1H MAS analysis indicatedthat the outer layer ofL-cysteine molecules interacting with the

    chemisorbed inner layer exhibit large amplitude motion aboutthe carboncarbon bonds.65 Comparison to L-cystine-function-alized AuNPs indicated that L-cysteine was not absorbed as thedisulde analogue. While these results are signicant forprotein-attachment strategies to AuNPs, their more importantimpact is in establishing REDOR as a powerful approach toresolve long-range interactions within the NMNP ligand shell,analogous to structural detail that has been transformative instructural biology.

    3.1.3. Ligand Shell Structural Dynamics. In addition toprobing the identity, quantity, and arrangement of ligands onmetal NP surfaces, NMR techniques are also useful to probevariations in structure (e.g., as a function of temperature) anddynamics of the NP ligand shell with atomic site resolution

    within the ligand.66For example, by simply measuring the 13C NMR chemical

    shift values of ligand resonances appended to a nanoparticle,the crystallinity of alkanethiol ligand shells can be estimated.This strategy takes advantage of 13C chemical shifts that areparticularly sensitive to trans- and gauche-conformationalchanges for ligands tethered to solid surfaces. Under MASconditions in the solid state, all-trans conformations for interiormethylenes resonate at 3334ppm and gauche conformationsappear between 2830 ppm.67 As expected, in solution adynamic equilibrium between trans and gauche conformationsis typically observed and leads to an averaged 13C chemical shiftof 2930 ppm, consistent with chemical shift averaging

    Figure 5.Ligand shell arrangements and morphologies can be assessedby combining solution phase 1H chemical shift with NOE analysis.(A), (B), and (C) represent the predicted chemical shift patterns as afunction of ligand composition. (D), (E), and (F) represent thepredicted NOE cross-peak patterns as a function of randomly mixed,

    Janus, and patchy ligand shell morphologies. Reproduced withpermission from ref62. Copyright 2012 Nature Publishing Group.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    H

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    9/19

    observed in other equilibrium processes monitored by NMR.More detailed information on alkyl chain dynamics, relativemolecular mobilities, and the degree of crystallinity can bemeasured with a combination of 13C and 1H T1 relaxationmeasurements,13C1H heteronuclear dipolar dephasing experi-ments (Figure 6B), and 2D 13C1H wide-line separationssNMR experiments.68

    Using the strategies above, chain ordering trends on AuNPs(d= 25nm) as a function of alkyl chain length have beenevaluated.68 In all cases, the thiol moiety is anchored to theAuNP surface, which is evident from NMR signal dephasing at13C positions adjacent to the thiol. When the alkanethiol chainlength is < C8, the ligand shells show dynamics similar to that ofthe solution phase (N.B. this does not imply rapid exchangewith solution phase ligands), indicating that a high populationof gauche conformers is present at short chain lengths. On theother hand, when the alkanethiol chain length is increased toC18, the ligand shell displays a high degree of conformationalorder from large-amplitude motion about the chain axes in a

    mostly trans conformation. The increased conformational orderwas proposed to be a result of chain intercalation in the solidstate with neighboring particles (Figure 6A).

    Surprisingly, ssNMR analyses also revealed a non-negligiblepopulation of gauche conformers concentrated at chain terminiin C18thiol-capped AuNPs, a detail that was not apparent fromFTIR analysis alone. While also a bulk measurement of theaverage, NMR has the distinct advantage of providing 13Cchemical shift resolution of specic carbon sites and differentelectronic environments along the alkane chain. Further, phasetransitions observed via DSC corresponded to reversible alkylchain disordering, as determined by variable temperature (VT)1H13C CP-MAS experiments. These results were later

    conrmed and expanded by Lennox and co-workers using acombination of DSC, FTIR, and 2H ssNMR.69

    In addition to crystallinity and chain ordering trends inalkanethiol-capped AuNPs, the inuence of terminal-alkylfunctional groups on these architectures was studied usingsimilar methods. Both conformational order and thermalstability have been studied for several chain lengths andfunctional groups including alcohols,68 carboxylic acids,70

    phosphonic acids,71 and sulfonic acids.71 In general, the authorsfound that hydrogen bonding imparted a higher degree ofconformational order and thermal stability compared tomethyl-terminated analogues.

    Solid state NMR techniques were used to resolve not onlyinteractions at the ligandsolvent interface but also to provideimportant insight into ligandparticle bonding at the hardsoftmatter interface. To describe in similar structural detail themetalsulfur binding motif present in AuNPs, Lennox, Reven,and co-workers investigated the AuSR interaction usingssNMR techniques for both long- (C14) and short-chain (C4)alkanethiols.72 Site-selective 13C isotopic labeling was used toenhance and resolve the 13C NMR resonances at positions C1and C2, closest to the thiol group. On the basis of chemicalshift comparison to the free ligands, the C1 and C2 positionswere both found to undergo extensive line broadening anddowneld shifts of 18 and 12 ppm, respectively. The changes inchemical shift upon coordination to AuNPs were consistentwith strongly adsorbed organosulfur species.

    Here, metal-specic NMR features were helpful in clarifyingthe origin of these changes in chemical shift. For example, theobserved chemical shifts could arise from coupling of the 13C1and 13C2 resonances to the conduction electrons on the AuNPsurface. However, the T1 values did not exhibit a linearrelationship with temperature. This lack of Korringa behaviorindicated that the resonances experienced little, if any, Knightshift contribution to the observed NMR chemical shift. Furthercomparison to analogous diamagnetic Au(I)thiolates showed

    similar changes in chemical shift at the C1 and C2 position tothose observed in the nanoparticle system. Taken together, theabsence of Korringa behavior and similarity to Au(I)thiolatechemical shifts, ruled out a metallic contribution and suggestedthe origin of chemical shift change was from the presence of aAuthiolate bond at the particle surface.

    Once the origin of the chemical shift difference was assigned,the source of line broadening was evaluated. The fwhm of boththe C1 and C2 resonances of the thiolate-capped AuNPs wereboth broadened signicantly (fwhm = 10001300 Hz). Asmentioned in Section 1.1, MAS technique s eliminatecontributions from CSA as well as isotropic bulk magneticsusceptibility sources of broadening. Upon MAS, the linewidths of C1 and C2 were reported to narrow only slightly,

    indicating that the contribution to the line broadening was theresult of a distribution of isotropic chemical shifts. A holeburning experiment conrms the heterogeneous line broad-ening of C2, which is likely the result of a chemical shiftdistribution from chemisorption of the thiol on variouscrystallographic sites of the AuNPs. These distributions inligand-particle bonding environments are then assigned to theimpact of a highly faceted particle surface (i.e., differentcrystallographic facets produce distinct environments for ligandand metal nuclei) as well as deviations in particle shape.Likewise, crystallographic variation in chemisorption sites wasalso independently found to cause heterogeneous line broad-ening in 31P and 1H resonances on triphenylphosphine-capped

    Figure 6. (A) Scheme of possible nanoparticle interactions andcorresponding TEM image of octadecanethiol-terminated AuNPs and(B) 13C1H dipolar dephasing measurements of the NPs shown in(A). (C) Correlation of static 2H NMR lineshapes with phasetransitions measured with DSC (Tm). (D) Scheme of trans and gauchemethylene confomers in equilibrium as measured by 2H NMRspectroscopy. (A) and (B) adapted with permission from ref 68.Copyright 1996 American Chemical Society. (C) and (D) adapted

    with permission from ref 69. Copyright 1997 American Chemical

    Society.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    I

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    10/19

    AuNPs (d = 1.8 nm), as determined by 31P and 1H hole-burning experiments.53 Both studies indicate that changes inline width and chemical shift upon particle attachment areprimarily the result of heterogeneous line broadeningmechanisms from a distribution of chemical environments onvarious surface facets.

    In both instances, NMR was able to provide unprecedentedinsight into the processes contributing to line broadening andchemical shift changes of the resonances closest to the ligandanchoring moiety on NMNPs. Further, there are notable casesin which NMR spectroscopy of the ligand shell may provideinformation about the chirality of the ligand arrangement oreven the underlying particle itself. Using a combination of 1Dand 2D solution phase 1H NMR techniques, Jin, Gil and co-workers demonstrated that glutathione-terminated Au25(SG)18clusters exhibit two different types of surface thiolate bindingmodes, consistent with previous NMR spectroscopy and X-raycrystallography results for Au25(SCH2CH2Ph)18.

    73 Alterationsin 1H chemical shift as a result of underlying Au nanoclusterchirality were also examined,74 similar to methods used intraditional organic synthesis. In both cases, the 1H NMRchemical shift behavior is sensitive to the surroundingelectronic environment, which includes the electronic structureand bonding environment of the nucleus. Therefore, changes inthe handedness of a molecule can be feltby neighboring spinpositions and is observed as a change in chemical shift, makingNMR observables valuable for assessing chirality or nonchiralityof small, molecule-like nanoclusters.

    In the case of small, noble metal clusters (

  • 7/25/2019 NMR of Nano Particles

    11/19

    on PtNPs followed Arrhenius behavior and extracted both anactivation energy and pre-exponential factor (6.0 0.4 kcal/mol and 1.1 0.6 108 cm2/s, respectively).84 The proposedmechanism for CO diffusion was exchange between differentCO populations driven by a chemical potential gradient. Thisstudy introduced a new method to quantitatively correlatediffusion of surface adsorbates to catalytic activity and should

    be amenable to the study of other adsorbed species and metalsurfaces.In a related study, the authors used a combination of 13CO

    electrochemical NMR and cyclic voltammetry (CV) toinvestigate the origin of Ru promoted electro-oxidation ofMeOH on PtRuNP catalysts.85 NMR results revealed twodifferent types of13CO on the catalysts, one population on purePt sites and another population on PtRu islands. Surprisingly,no exchange was observed between different CO populations,and these observations were supported by CV measurements.Detailed analysis of 13C chemical shift changes and Korringabehavior between the two CO populations suggested that Ruweakens the PtCO bond, resulting in increased CO oxidationrates. The combination of NMR spectroscopy and electro-

    chemistry provided unprecedented insight into CO toleranceand promotion in bimetallic NP catalysts.3.3. Metal Nanoparticle Core Characterization.

    3.3.1. Nanoparticle Size. A basic property of any nanoparticleis its size. NMR is a useful tool to measure the hydrodynamicradius of metal nanoparticles and can provide an importantcomplement to traditional nanoparticle sizing techniques, suchas electron microscopy and DLS. Similar to DLS, NMR signalcan be used to determine nanoparticle size via analysis ofparticle diffusion. Specically, NMR uses pulsed-eld gradient(PFG) techniques to extract diffusion coefficients of well-dispersed species in solution diffusing according to Brownianmotion only. Under these conditions, the hydrodynamic size iscalculated by rearranging the StokesEinstein equation:

    =D

    k T

    R6

    B

    H (3)

    where D is the diffusion coefficient, kB is the Boltzmannconstant,Tis temperature, is viscosity of the solvent, and RHis the hydrodynamic radius of the diffusing species.

    However, unlike DLS, NMR provides the added benet ofchemical resolution. In 1H (or other nuclei of interest) DOSY,a pseudo-2D NMR experiment is performed that separatesNMR signals according to their diffusion coefficient. Murrayand co-workers presented one of the rst accounts of1HDOSYto measure the hydrodynamic size of Au colloids86 andextended this approach to systematically study the size

    dependent 1

    H DOSY signatures of AuNPs with core sizesranging from 1.5 to 5.2 nm.51 Here, the authors also noted acorrelation between nanoparticle core size and the fwhm of the1H NMR spectra, which was a result of the slower moleculartumbling and reduced correlation time (and, hence, decreasedT2 values) for larger particle sizes. Using changes in NMRspectral breadth as a function of particle size may be particularlyuseful for systems that are not amenable for DOSY analysis,such as dendrimer-encapsulated nanoparticles.87 It is worthnoting that DOSY analysis can be combined with NOEtechniques to use NMNPs as a platform for NMR-basedchemosensing of small molecules in solution as reported byMancin and co-workers.88,89

    Building upon earlier work, Kubiak and co-workers reporteda convenient nanoparticle sizing technique using 1H DOSYmeasurements in deuterated organic solvents.90 The sizesmeasured with DOSY compared well with results obtainedfrom TEM for AuNPs ranging in size from d = 25 nm.Nanoparticle size distributions were extracted from the datausing the continuous method CONTIN.91 An importantcontribution of this study is the use of an internal standardfor the calibration of nanoparticle sizing. The use of an internalstandard mitigates experimental error while simultaneouslysimplifying data processing. Here, an internal standard withboth a wel l-dened hydrodynamic size and 1H NMRresonances well-resolved from that of the nanoparticle ligandresonances is added to the sample before 1H DOSYmeasurement. Common examples of internal standards includeferrocene for organic solvents and dioxane for aqueous media.During the experiment, the diffusion coefficients of both theinternal standard and the nanoparticle-bound ligands aremeasured. The hydrodynamic radius of the nanoparticle canthen be calculated from eq4:

    =R

    D

    D RNPref

    NPref (4)

    where RNP is the hydrodynamic radius of the nanoparticle, Drefis the diffusion coefficient of the reference molecule asmeasured by 1H DOSY, DNP is the diffusion coefficient of theligands bound to the nanoparticle as measured by 1H DOSY,and Rref is the known hydrodynamic radius of the referencemolecule in the solvent of interest. The measurement of relativediffusion coefficients minimizes error from changes in sampletemperature, viscosity, instruments, independent measure-ments, and uctuations during data acquisition. In ourlaboratory, we have found that this method can be extendedto measure the hydrodynamic size of a range of metalnanoparticle compositions in polar and nonpolar solvents,

    and that sizes obtained match well with those measured byTEM.92,93

    Further, we have also found that, by changing the chainlength of capping ligand on the nanoparticle surface, we areable to tune and detect corresponding changes in hydro-dynamic radius using 1H DOSY. Likewise, Hakkinen and co-workers found that the apparent diffusion coefficient andhydrodynamic size of Au nanoclusters (144 and 102 atoms)capped with p-merpcaptobenzoic acid (pMBA) in aqueoussolution depends strongly on the counterion of thedeprotonated pMBA capping ligand.94 DFT calculationsrevealed that competing hydrogen bonding interactions andion-pairing between the pMBA, Na+, NH4

    +, acetic acid, andwater molecules affected the hydrodynamic size of the Au

    nanoclusters.Despite these successes, measuring nanoparticle size

    becomes more challenging and time-consuming with largerparticle diameters due to the NMR line broadening effectsobserved from slower tumbling at larger sizes. It may be that 1HDOSY is most useful at size ranges where alternate sizingtechniques, such as TEM imaging, becomes less useful (due toboth instrument and sample limitations). It is also important tonote that NMR is a population-averaged technique, meaningthat 1H DOSY could introduce a bias toward larger particlesizes when performing sizing measurements, assuming thatlarger diameter particles contain more capping ligands thansmaller ones. However, with these caveats in mind, the

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    K

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    12/19

    combination of 1H DOSY, TEM, and UVvis can providerobust and sometimes otherwise inaccessible information onthe average size of metal nanoparticles.

    3.3.2. Nanoparticle Core Properties Observed Using MetalNuclei. Here, we discuss NMR spectroscopy of metal nucleicontained within the nanoparticle itself. These experimentsreveal detail both about the nanoparticle physical architectureas well as its electronic structure. For a comprehensivediscussion on the theory and history of metal NMR to studyparticles and clusters, we refer the readers to a classic review byvan der Klink and Brom from 2000.5 Likewise, the quantum dotand metal oxide communities have leveraged direct NMRobservation of nuclei that compose the nanomaterial to learnabout details such as surface architecture95 and particleelectronic structure96 a s a f unction of size. For metal-

    s,109Ag,97101 103Rh,102 and 63Cu103106 NMR of small metalnanoparticles have all been reported, but 195Pt is the most well-studied nucleus, because of favorable NMR properties such asrelatively high natural abundance (33.8%) and moderategyromagnetic ratio. Here, we focus on reports of 195Pt NMRwith the note that techniques outlined for Pt may be extendedto other NMR active metal nuclei as both instrumentation andmethodology continue to advance.

    Although 195Pt NMR had been used previously to study arange of PtNP systems,107 Slichter and co-workers reported oneof the rst observations of size dependent spectral changes in195 Pt NMR line shape as a funct ion of metal corediameter.4,108,109 This exciting result was conrmed by vander Klink and co-workers who also reported a broad 195Pt

    NMR line shape spanning 2.5 MHz at 8.5 T from sur facetocore resonances in the presence of additional adsorbates.110112

    Ab initio calculations suggested that the 195Pt NMR shift in thesurface Pt species compared to the core was the result of agradual decrease in the d-like Fermi-level local density of states(Ef-LDOS) upon moving from the inside of the particle to itssurface.113 Further, Slichter and co-workers found that 195PtNMR lineshape was a function of particle size and adsorbateidentity.108 As particle size decreased, the 195Pt NMR peakcorresponding to bulk Pt metal also decreased. Likewise, whenthe particles were coated with adsorbates, a surface peak wasobserved in the region where diamagnetic 195Pt species aretypically observed. When the surface of the particles was

    cleaned (heat treated to remove adsorbed molecules), thepeak disappeared. Additionally, the frequency of the 195Pt NMRsurface peak was dependent upon the chemical identity of theadsorbate.

    In order to extract more quantitative information from wide-line static 195Pt NMR lineshapes, Bucher and co-workersdeveloped a layer model,110,114 which assumes a pseudospher-ical particle shape, and each population-weighted layer of Ptatoms contributes to a specic NMR frequency (Figure7A).115

    In this model, the average Knight shift of a given layer, n, can bewritten as follows:

    = +

    K K K K( )e

    n

    n m

    0( / )

    (5)

    where K0 is the Knight shift of the surface, K is the Knightshift of bulk Pt, and mis the healing length. Here, the healinglength is a probe of how strongly the more molecule-likesurface is able to inuence the metallic Knight shift of theinterior Pt atoms. The healing length will vary with particlediameter, but Oldeld, Wieckowski, and co-workers havedemonstrated that healing length can also depend on theelectronegativity of adsorbates (vide infra).116

    The difference in healing length can be seen in the shift ofNMR frequency as a function of PtNP core diameter (Figure7B). As the population of surface atoms increases, thepopulation of surface species present in the spectrum increasesaccordingly. Likewise, to examine the inuence of Pt corediameter (d 15 nm) on catalytic behavior, Watanabe,Oldeld, Wieckowski, and co-workers used 195Pt NMR analysis

    to study the oxygen reduction reaction (ORR) in anelectrochemical environment.117 Surprisingly, 195Pt NMR lineshape andT1relaxation analysis indicated that surface Pt atomsshowed similar electronic structure, regardless of core size, andthus had negligible effect on ORR rate constants. This workindicated that the electronic structure of the surface Pt speciesalone may dictate catalytic behavior, rather than changes inparticle size (it is important to note that at NP sizes where thenumber of surface atoms dominates the total atom populationof the particle, these two effects (core size vs surface structure)may be indistinguishable).118

    In addition to the inuence of particle size on Ef-LDOS, theinuence of particle surface chemistry has also been examined

    Figure 7.(A) Wideline 195Pt NMR spectrum of 5 nm carbon-supported PtNPs showing deconvolution as a function of atom position. (B) Changein 195Pt NMR line shape as a function of PtNP core diameter and % surface atoms. Adapted with permission from ref115. Copyright 2013 RoyalSociety of Chemistry.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    L

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    13/19

    using 195Pt NMR. For example, comparison of 195Pt NMRlineshapes, chemical shifts, and T1relaxation behavior of clean,K-, and Li-impregnated PtNP catalysts revealed that alkali salt-impregnation increased Ef-LDOS at surface Pt sites by 1015%.119 This result was in contrast to H2adsorption, which wasshown to diminish Ef-LDOS at Pt surface sites. The spatialresolution available with 195Pt NMR analysis also indicated thatthe alkali salt adds to the surface of the PtNPs and does notdiffuse to the particle interior. These data provided a structuralbasis for proposed mechanisms of alkali salt promotion in PtNPcatalysts.

    Holding PtNP core size constant (d 2.5 nm), Oldeld,Wieckowski, and co-workers systematically investigated theinuence of a variety of adsorbates (H, O, S, CN, CO, andRu) on 195Pt NMRgures of merit.116 Here, the authors foundthat the Knight shift of interior Pt atoms remained unchanged,regardless of surface chemistry. However, the Knight shift ofthe surface and subsurface Pt sites varied over 11 000 ppmand showed an increasing downeld shift as the electro-negativity of the adsorbate increased. Similarly, in twoindependent reports, Pt atoms bound to organic cappingligands exhibited no Knight shift and did not show typicalKorringa behavior, while Pt atoms at the core exhibited metallicproperties.120,121

    Another class of relevant Pt-containing nanostructures thatare particularly difficult to characterize are small multimetallicarchitectures (d= 13 nm) containing Pt and at least one othermetal. In these materials, the number as well as the position ofthe constituent metal atoms are crucial to their optical, catalytic,and magnetic properties. Methods such as X-ray absorptiontechniques as well as ssNMR can provide comprehensiveinformation on metal nanoparticle architectures. NMR has theadvantage of being more accessible and therefore can providereal-time analysis of ongoing experiments. Especially in the caseof 195Pt NMR, the electronic structure at both the surface andcore of the nanoparticle can be determined by direct

    observation of the metal nucleus as described in pure Pt NPsabove.

    One of the rst reports of 195Pt NMR to probe bimetallicnanoparticles investigated PtRh NPs and correlated theresults with metal segregation observed in EDS.122 Furtherinvestigations used 195Pt NMR to understand the changes inEf-LDOS of catalytically active PVP-coated PtPd nanoparticles.Here, van der Klink and co-workers showed that the Ef-LDOSof the interior Pt atoms varied strongly with % Ptcomposition,similar to observations in bulk PtPd alloys.123 The authorssuggested that the changes in Ef-LDOS were also present at thesurface of the NPs and were responsible for changes in catalyticbehavior as a function of composition in PtPd particles.

    In a separate report, Oldeld, Wieckowski, and co-workers

    used a combination of TEM, electrochemistry, and 195Pt NMRspectroscopy to investigate the effect of heat treatment onbimetallic PtRu nanoparticle alloys (d= 23 nm).85,124 Uponheat treatment at 600 C, the 195Pt NMR signal of PtRu NPsshifted upeld, consistent with Pt migration to the interior ofthe particle.85 This structural change resulted in a decrease incatalytic activity and was also consistent with a decrease in thenumber of Pt atoms on the surface of the NP. When theoriginal sample was instead subjected to heat treatment at 220C in H2(g), CO tolerance and methanol oxidation reactivityboth increased, consistent with increased metallic Ru at theparticle surface. At the same time, 195Pt NMR resonancesexhibited a change in Korringa product,T1T, consistent with a

    decrease inEf-LDOS of the Pt. Taken together, the changes incatalytic behavior and 195Pt NMR properties led to the proposalof a Ru-rich core with a PtRu alloy overlayer as a result ofheat treatment at 220 C in H2(g). Later, Tong and co-workersexpanded this approach to include 195Pt NMR shift analysisforspatial distribution of Pt atoms in PtRu nanoparticle alloys125

    as well as to probe electronic structure in other bimetalliccompositions, including PtAu NPs.126

    Recently, Hanna and co-workers investigated a series of Pt3X(where X = Sn, Al, Sc, Nb, Ti, Hf, and Zr) bimetallicnanoparticles with 195Pt NMR spectroscopy.115 Here, theauthors also used ssNMR analysis of the heteronuclei present inthe Pt3X alloys, which provided multielement informationcomparable to powder X-ray diffraction (XRD) techniques(XRD was also performed and correlated to the NMR results).Multinuclear NMR comparison with XRD facilitated assess-ment of bimetallic NP composition, size, relative order/disorder, and electronic structure. This study highlights theinsight that can be achieved by combining NMR techniqueswith traditional materials characterization tools.

    Despite its broad utility, routine acquisition of wide-lineNMR spectra can be technically demanding. The inherentdifficulty in acquiring 195Pt NMR of Pt-containing nanoparticleslies in the fact that the static lineshapes often span severalmegahertz, making uniform broadband excitation challenging.Several approaches have been used to reconstruct ultra-wide-line patterns, including spin echo height spectroscopy (SEHS),variable offset cumulative spectroscopy (VOCS),127 and eldsweep Fourier transform (FSFT) spectroscopy.115 Thedevelopment of methods such as wideband uniform ratesmooth truncationCarrPurcell MeiboomGill (WURST-CPMG)128promise to greatly reduce the time and sensitivityburden associated with collecting ultra-wide-line spectra, whileaccurately replicating lineshapes. Importantly, the advent ofbroadband excitation and sensitivity enhancement techniques,such as FSFT and WURST-CPMG, suggest the opportunity to

    explore more exotic metal elements such as 105Pd and 197AuNMR as well as dramatically expand the characterization of Pt-containing NP systems.

    3.3.3. Nanoparticle Core Properties Observed via Adsor-bate Nuclei. Above, we describe the use of NMR to directlyobserve nuclei in the nanoparticle core. However, NMR of theadsorbate may also provide information about the morphologyand electronic structure of the NP. For example, certainadsorbates such as 13CO exhibit a Knight shi ft andcorresponding Korringa behavior as a result of mixing betweenthe adsorbate molecular orbitals and the transition metal dband.4

    Slichter and co-workers conducted a large body of founda-tional work examining adsorbates on the surface of metal

    particles.4,129 In 1985, 13CO adsorption on PtNPs was found toshift the 13C resonance to much higher frequency (200 ppmfrom the unbound CO resonance).130 The authors suggestedthat this large shift was a Knight shift, and therefore the resultof polarization of electron spins. This assignment wassupported by the observation of Korringa behavior, andspecically the temperature dependent T1 of the

    13COmolecule. Additionally, the Ef-LDOS on the C atom wasdetermined from electron spin resonance measurements onCO radicals. Slichter and co-workers went on to measure the T1behavior of 13CO adsorbed on a variety of small metal particlecompositions including Ru, Pd, Rh, Os, and Ir and were able todraw similar conclusions.4

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    M

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    14/19

    A quantitative correlation between the Knight shift ofchemisorbed 13CO and the Ef-LDOS on the surface of Ptand Pd nanoparticles was the subject of a later report byOldeld, Wieckowski, and co-workers.131 In this work, theauthors investigate the NMR properties of 13CO adsorbed toPtNP and PdNP catalysts in an electrochemical environment.The data compared the 13C Knight shift,Ef-LDOS values fromDFT calculations and NMR measurements of four systems:13CO adsorbed onto M7CO clusters (where M = Pt or Pd),

    131

    13CO adsorbed onto oxide-supported PtNPs in a dryenvironment,4,110 13CO adsorbed onto carbon-supportedPtNPs in a wet electrochemical environment,131 and 13COadsorbed onto oxide-supported PdNPs in a dry environment(Figure 8).132,133 The linear relationship in Figure 8 was

    supported by trends in nanoparticle size as well as IRmeasurements of particle-bound CO.131,134 The authorssuggested that the linear correlation between the Knight shiftof chemisorbed CO, K13CO, and the clean Ef-LDOS of thetransition metal substrate followed the form:

    = K a LDOS13CO (6)

    where a 11 (2) ppm/Ry1 atom1 for Pt and Pd.While observation of the 195Pt NMR resonance from the

    PtNP catalyst can serve as a direct probe of theEf-LDOS of themetal nanoparticle, direct NMR observation of metals such as105Pd nuclei remains challenging. However, 13CO may be usedas a probe of the underlying PdNP. Specically, the authors

    measured the Knight shift and T1 values of linear and bridge13CO adsorbed onto the surface of PdNPs as a function oftemperature.132 These results are important because theysuggest that the NMR behavior of the ligand nuclei bound to ametal surface can be used to probe the metal electronicproperties, even in the case of NMR-silent metals. Thecorrelation of ligand nuclei with metal core electronicproperties dramatically expands the utility of NMR tounderstand the behavior of catalytic and photoactive nanoma-terials.

    Both theoretical135 and experimental NMR6 studies provideevidence that, indeed, adsorbates other than 13CO can probethe metallic and structural properties of nanoparticles. For

    example, Kitagawa and co-workers used 2H2(g) to examinemorphology changes in bimetallic NPs. In these studies, theauthors used 2H NMR spectroscopy to distinguish betweencoreshell and alloyed architectures of PdPt,136 PdAu,137

    and AgRh138 nanoparticles, morphologies which could betuned as a function of atomic composition.

    Tong and co-workers used 77Se NMR to examine theinuence of selenol-terminated ligands on underlying AuNPelectronic properties. In this case, results indicated changes inboth the chemical shift and the temperature dependent T1relaxation rate of the Se nucleus, consistent with strongcoupling to electrons on the Au surface.139 The NMR behaviorobserved in this case is consistent with a possible Knight shiftcontribution to the 77Se NMR resonance. Unfortunately,analogous 33S NMR experiments are not suitable for routineassessment of thiol-binding environments because 33S exhibitsunfavorable NMR properties such as a moderate quadrupolemoment (6.78 1030 m2), low natural abundance (0.76%),and low gyromagnetic ratio (2.06 107 rad T1 s1). However,based on the work presented by Tong and co-workers, 77SeNMR of selenol-capped nanoparticles is a promising alternativeprobe of the NMNP metallic properties.

    4. USING NMR TO ASSESS NANOPARTICLEPERFORMANCE

    Because of its broad accessibility and its ability to analyzeNMNPs in situ, NMR may be uniquely well-suited to monitorNMNP performance in certain applications. Here, we giveexamples of how NMR spectroscopy can provide robuststructurefunction correlations between NMNP architectureand its utility in applications such as bioimaging andheterogeneous catalysis.

    4.1. Magnetic Properties.NMR techniques can be used tomeasure gures of merit for magnetic resonance imaging(MRI) contrast agents. The principles that apply to metalnanoparticle systems are identical to other contrast agents andare not discussed in detail here. Briey, in order to ascertain theefficacy of an MRI contrast agent, T1and T2relaxation rates ofsurrounding media are measured as a function of metal ornanoparticle concentration using standard inversionrecoveryand CPMG pulse sequences, respectively. The slope of theresulting relationship between T1 or T2 relaxation rate vsconcentration is referred to as the relaxivity value, r1 or r2,respectively, which can be used to compare between differentcontrast agents.

    In addition to MRI, NMR measurements can provide afundamental understanding of the magnetic susceptibility incolloids. The NMR-based Evans method140 is a particularlyattractive technique to measure magnetic susceptibility becauseit can be used to rapidly evaluate the magnetic properties of

    NPs in solution (the entire colloid is analyzed), it does notrequire a large amount of material (1 nmol NPs), and it is arelatively simple, widely accessible procedure (1D 1H NMRacquisition, performed on any NMR instrument). Importantly,the Evans method is also more tolerant of surroundingdiamagnetic materials appended to the NP surface (e.g., smallmolecule ligands), in contrast to other susceptibility measure-ments such as superconducting interference quantum device(SQUID) measurements.

    In the Evans method, the mass magnetic susceptibility ismeasured by comparing the 1H NMR chemical shift of asolution containing the magnetic colloid and a standard withthat of the pure standard (if suitable, the solvent can serve as

    Figure 8. 13CO Knight shift as a function of clean surface LDOSbased on ab initio calculations of CO adsorbed on Pt7clusters (openblack square) and experimental NMR measurements of 13COadsorbed on Pt (closed black squares) and Pd (closed blue circle)nanoparticle substrates. Adapted with permission from ref 131.Copyright 1999 American Chemical Society.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    N

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    15/19

    the standard). The comparison can be made using a coaxialinsert inside an NMR tube, with the inner tube containing purestandard and the outer tube containing both the standard andthe colloid of interest. This experimental approach allows

    acquisition of both species in a single 1D 1H NMR spectrum.The difference in 1H NMR frequency between the standard inthe colloidal solution and the pure standard is related to thetotal mass susceptibility by a modied Evans methodequation:141

    =

    +f

    fm

    3

    4tot,g 0 (7)

    Here, tot,g is the total mass susceptibility, fis the frequencydifference in Hz, f is the operating frequency of thespectrometer, m is the mass of magnetic species in 1 mL ofsolvent, and 0 is the mass susceptibility of the solvent. TheEvans method has been used to rapidly assess the magneticproperties of superparamagnetic metal oxides and inorganiccomplexes and has started to be used to evaluate thesusceptibility of binary metal nanoparticles, including AuNi142 and AuCo.93

    Chandler and co-workers used the Evans method todetermine the room-temperature, solution-phase magneticsusceptibility of AuNi nanoparticles (d = 3 nm) and alsomeasured these values using SQUID measurements.142 In thisstudy, temperature-dependent SQUID measurements >10 Kwere difficult to obtain because of the diamagnetic contributionfrom excess dendrimer template, alkanethiols, and residualsolvent necessary to stabilize the bimetallic structures. Here, theEvans method provided an alternative to measure the netmagnetic susceptibility at room temperature without perturbingsample integrity.

    Recently, we used the Evans method to determine thecomposition-tunable magnetic properties of AuxCoyNPs as afunction of % Co incorporated in the nal NP.93 By increasingthe concentration of Co in the nal NP, the mass susceptibilityof the NPs was tunable from 3.9 107 to 112.6107 cm3/g (Figure 9). All measurements were performed at roomtemperature, in D2O, with nanoparticle quantities of1 nmolNPs. Spectral acquisition was typically complete within 30 s,allowing high throughput analysis of many samples andcompositions with little demand on material quantity.

    In the previously described analysis, the total masssusceptibility is comprised of both the diamagnetic andparamagnetic contribution. If the molecular weight, M, of the

    nanoparticle is known, the paramagnetic contribution can beextracted by converting the mass susceptibility to molarsusceptibility according to eq8and subtracting the diamagneticcontribution using Pascals constants143 (eq9).

    = Mtot,mol tot,g (8)

    = para, mol tot, mol dia, mol (9)

    = T8effpara,mol (10)

    From here, the effective magnetic moment, eff, pernanoparticle can be calculated according to eq 10, usingelemental analysis of particle concentrations. However, guresof merit such as blocking temperature, saturation magnet-ization, and extent of hysteresis must be determined usingalternate magnetic characterization techniques such as SQUID.Yet, because SQUID is technically more demanding thanNMR, NMR measurements of magnetic properties are anattractive complement and/or alternative for many NMNPinvestigations, including rapid screening in materials develop-ment for applications such as data storage, bioimaging, andsupercomputing.

    4.2. Catalytic Behavior. In addition to providing rapid,high throughput information on the magnetic properties ofNMNPs, NMR can also be used to explore catalytic behavior.Several examples in the literature use NMR to monitorreactions on metal nanoparticle catalysts, some of which werehighlighted in Section 3.2. Recently, Tsang and co-workershave used NMR to monitor reactant turnover in heterogeneouscatalysis, but have also correlated changes in 13C chemical shiftof chemisorbed formic acid with catalytic gures of merit.Specically, the authors found that 13C chemical shift of

    adsorbed formic acid was related to the work function of thesurface and the specic activity for a variety of carbonsupported and PVP-coated colloids with both monometallicand bimetallic core@shell compositions (Figure 10).144

    Previously, the authors provided evidence that 13C-labeledformic acidcan probe the electronic properties of Ru particlesurfaces.145 In this study, the authors introduce an oxygenspacer between the 13C label and the particle surface and wereable to eliminate the line broadening from Knight shift effects(e.g., those observed with 13CO probe molecules). Thisapproach was extended to several other monometallic andbimetallic NP systems. For 13C-labeled formic acid adsorbed toPVP-coated PdNPs, four separate resonances were observed

    Figure 9.(A) Evansmethod 1H NMR spectra of the HDO peak in pure D2O (asterisk) and the HDO peak in D2O containing various colloidalcompositions. (B) Mass susceptibility as a function of % Co incorporated in AuxCoyNP alloys, as measured from the Evansmethod spectra in (A).

    Yellow circles = Au atoms. Blue circles = Co atoms. Adapted with permission from ref93. Copyright 2014 Wiley-VCH Verlag GmbH & Co. KGaA.

    Chemistry of Materials Perspective

    DOI: 10.1021/cm504809cChem. Mater. XXXX, XXX, XXXXXX

    O

    http://dx.doi.org/10.1021/cm504809chttp://dx.doi.org/10.1021/cm504809c
  • 7/25/2019 NMR of Nano Particles

    16/19

    (Figure10A). The peak at 165.89 ppm was assigned to weakly

    adsorbed formic acid that was in rapid exchange with freeformic acid, leading to an average chemical shift. The remaining13C resonances, 165.42, 165.69, and 165.95 ppm, were assignedto monodendate, multimonodentate (see Figure 10A formolecular structure), and bridging formate adsorbed to theparticle surface, respectively. 13C NMR spectral assignment wascorrelated with FTIR spectroscopy results.

    Remarkably, even in the presence of an oxygen spacer, DFTcalculations suggested a signicant degree of overlap of the 2sand 2p orbitals of the 13C atom with metal d-electrons whenformate was in a bridging conformation. Linear trends wereobserved between 13C chemical shift of the bridging formate onthe NP surface and the d-band center, work function (Figure10B), and specic activity (Figure 10C) of the underlyingcatalysts. Similar trends were observed for a variety of metaltypes, particle sizes, and compositions, emphasizing the breadthof information gained by using NMR spectroscopy for routineNP performance evaluation.

    5. OUTLOOK

    From the work highlighted in this Perspective, it is clear thatNMR spectroscopy has broad utility in the eld of NMNPs,both in terms of the NMR techniques available and the NMNPproperties measured.

    Yet, it is also clear that there are challenges to capitalizing onthis utility. The rst barrier is low and largely logistical: manyresearchers trained and participating in nanoparticle studies arenot also trained in the use of NMR as an analytical tool and

    likewise many experts in NMR spectroscopy are not active innanomaterials research. This is changing rapidly as the utility ofNMR in day-to-day nanochemistry work (e.g., particle sizing orligand shell characterization) becomes apparent. Second, asmentioned previously, all NMR studies suffer from the inherentlow sensitivity of Zeeman splitting. There are ongoing efforts toovercome this disadvantage with high eld instrumentation,specialized hardware and pulse sequences,146,147 and techniquessuch as dynamic nuclear polarization.148 Many of thesestrategies have already been used successfully in the eld ofstructural biology, and it is likely that these advances can also beapplied to the study of NMNPs. Finally, beyond basic 1 and 2DNMR techniques, both the technical and conceptual challenges

    of NMR spectroscopy increase steeply. For example, experi-

    ments that combine electrochemistry and NMR within a singleinstrument are beyond the capabilities of many but the mostexpert NMR researchers. In these cases, we hope thisPerspective will encourage expanded collaboration betweenthe two disciplines so that, together, we can ask the mostpressing scientic questions and answer them using the mostaccurate and efficient tools available.

    Taken together, the work presented here demonstrates thatNMR spectroscopy is a powerful complement to andsometimes an advantage over traditional materials character-ization techniques for NMNPs. With a long history to buildfrom, the future of NMR spectroscopy in the study of thesesystems is bright and should yield transformational insights intoNMNP formation, structure, and performance.

    AUTHOR INFORMATIONCorresponding Author*E-mail:[email protected].

    NotesThe authors declare no competing nancial interest.Biographies

    Lauren E. Marbella earned her B.S. in biochemistry from DuquesneUniversity in 2009, where she performed undergraduate researchunder the direction of Partha Basu. She earned her M.S. at theUniversity of Pittsburgh under the direction of Megan M. Spence,studying peptide binding and morphology changes in lipid membranesusing solid state NMR. In 2012, she joined the laboratory of Jill E.Millstone as a Ph.D. candidate investigating the formation and nal

    particle characteristics of noble metal nanoparticles using magneticresonance techniques.

    Jill E. Millstone received her B.S. in Chemistry and English fromCarnegie Mellon University in 2003, where she carried outundergraduate research in the laboratory of Richard D. McCullough.She completed her Ph.D. in materials chemistry in 2008 atNorthwestern Univer