synthesis, characterization, and reactivity of …

195
SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF VOLATILE COMPOUNDS FOR MATERIALS APPLICATIONS BY BRIAN J. BELLOTT DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemistry in the Graduate College of the University of Illinois at Urbana-Champaign, 2010 Urbana, Illinois Doctoral Committee: Professor Gregory S. Girolami, Chair Professor Patricia A Shapley Professor M. Christina White Professor Catherine J. Murphy

Upload: others

Post on 24-Dec-2021

12 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF VOLATILE COMPOUNDS FOR MATERIALS APPLICATIONS

BY

BRIAN J. BELLOTT

DISSERTATION

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemistry

in the Graduate College of the University of Illinois at Urbana-Champaign, 2010

Urbana, Illinois

Doctoral Committee:

Professor Gregory S. Girolami, Chair Professor Patricia A Shapley Professor M. Christina White Professor Catherine J. Murphy

Page 2: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

ii

Abstract

Elemental boron is a highly attractive fuel for propellants and explosives, but in order to

increase the rate of energy release during the combustion of boron and make it a more attractive

fuel for propellants and explosives, it is advantageous to prepare it in a nanoparticlulate form.

Boron nanoparticles were prepared by gas phase pyrolysis of decaborane. The particles are >97

% boron and free of hydrogen as judged by combustion analysis. Typical recovered yields are

40 %. In order to incorporate nanoparticles into larger assemblies and to stabilize them for long

term storage under ambient conditions, it is desirable to functionalize their surfaces. Treatment

of the boron nanoparticles with a ~6 mol percent benzene solution of XeF2 (a convenient

fluorinating agent), yields boron particles whose surfaces are fluoride terminated.

Thermogravimetric analyses (TGA) conducted in pure oxygen show that surface fluorination of

the boron nanoparticle causes an increase in the onset temperature of oxidation and increases the

overall yield of oxidation products.

High energy nanomaterials such as nanothermites and Al nanoparticles have been a topic

of increasing research for applications as propellants and explosives. It is well known that a

strong correlation often exists between the energy density of a nanomaterial and its sensitivity; a

related issue is that nanoparticles are strongly driven to agglomerate and densify, owing to their

relatively high surface free energies. One approach to addressing these problems is to passivate

the nanomaterials by functionalizing their surfaces. In 2005, Higa and co-workers claimed that

treatment of aluminum particles with solutions of transition metal acetylacetonate (acac)

complexes resulted in the deposition of a uniform coat of the zerovalent transition metal that

stabilized the nanoparticle toward oxidation. We find that treatment of aluminum nanoparticles

Page 3: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

iii

with transition metal acac complexes yields aluminum particles decorated with discrete transition

metal nanocrystals and does not result in a core-shell structure. Powder X-ray diffraction and

scanning transmission electron microscopy (STEM) analysis of the particles reveals the

nanocrystals vary in size depending on the transition metal. For platinum a narrow size

distribution of small nanocrystals (1-3 nm) decorate the surface of the aluminum particles

whereas for gold a broad size distribution of nanocrystals (2-70 nm) decorate the surface.

In order to explore the preparation of new types of copper alkene complexes for CVD

and ALD applications, we have investigated the structures and metal-alkene binding studies of

copper(I) triflate of three chelating tri-alkenes: the cis,cis,trans (cct), cis,trans,trans (ctt), and

trans,trans,trans (ttt) isomers of 1,5,9-cyclododecatriene (cdt). The X-ray crystal structures show

that, in all three compounds, the triflate ligand and all three C=C double bonds of the triene are

bound to the copper center, which adopts a distorted tetrahedral coordination geometry.

Competitive binding studies in toluene show that free energy of the copper ctt complex is lowest,

and that the free energies of the ttt and cct isomers are higher by 0.3 and 3.7 ± 0.1 kcal mol-1,

respectively. For the ttt and cct isomers, binding to the metal is attended by conformational

changes that increase the conformational energy of the ring; this increase is a penalty that

destabilizes the resulting copper complexes relative to that of the ctt isomer.

Page 4: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

iv

To all those I love; Thank you for everything.

Page 5: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

v

Acknowledgements:

I want to start off by thanking my family. Mom and Dad thank you for always

supporting me in the brightest and darkest of times, your unwavering support is my foundation

and I cannot express how much you mean to me. To my brother Trent and his wife Amanda,

thank you for the fishing trips, dinners, and occasional odds and ends. Trent your support of me

through this degree has meant more to me than words can describe. You never once doubted my

ability or drive and I thank you for always taking care of me.

Paul Sarazin thank you for teaching me the importance of factor label notation and

motivating me to set goals higher than I thought I could reach. Thanks to Richard and Ellen

Keiter, who showed me how conduct chemical research and motivated me to apply to graduate

school. I will always remember our games of checkers, moving rail road ties, lunches, and

dinners. Richard I cannot thank you enough for offering me my first lab position all those years

ago. I am glad you saw in me the ability to conduct research, it has become one of my greatest

joys in life. Thanks to Greg Girolami, I appreciate all the time and hard work you have put into

me. I hope to continue our collaboration for many years to come. Thanks to Vera Mainz for

hosting so many memorable and enjoyable Girolami group outings. You have showed me that

one person can tackle hundreds of obstacles without breaking a sweat. Thank you to my

committee members Christina White, Pat Shapely, Andrew Gewirth, and Cathy Murphy. I

appreciate all the help and criticism you have given me thus far.

Thanks to past and present Girolami Group members Amy Whelpley, Charity Flener,

Wontae Noh, Andrew Dunbar, Chuck Spicer, Paul Dickinson, Richard Jew, Do Young Kim,

Andrew Sealey, Justin Mallek, Noel Chang, Jennifer Steele, Luke Davis, Charles Heidbreder,

Mark Nesbit, Jessie Creamean, Ben Suslick and Collin Brandt. You all have helped me in so

Page 6: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

vi

many ways from listening to my ideas to aiding me in the discovery of new chemistry thank you

for everything. Chuck thanks for joining the lunch crew, our many conversations about sports,

life and chemistry will always be fond memories. Paul thank you for helping me out on those

many nights early in my career, your ability to understand and teach is greatly appreciated.

Richard thank you for always keeping me in check and proofing whatever I was writing, you are

an excellent scientist and I hope you pursue a tenure position in the near future. Andrew Sealey

thank you for showing me that there is more to life than lab work, I will always remember our

many conversations about life, football (yours and mine), billards, and England. Luke thank you

for allowing me to “borrow” your hood the last few months of my degree, without the extra hood

I am not sure how I would of made all the new compounds I did. Andrew thanks for always

having the screw driver ready! Finally I would like to thank Scott Daly you have become an

amazing chemist right before my eyes, thank you for all your help, time and understanding.

Your ability handle your career and personal life is truly amazing. I look forward to our

continued friendship and collaboration.

Thanks to all my good friends I have made while in Champaign-Urbana Nora Wang,

Ezra Eibergen, Zach Heiden, Brad Gorecki, Carlisle D’Souza, Brandon Lange, Chris Field,

Brittany O’Connor, Chris Letko, Jessica Herzog, Keith Porter, Luke Thompson, Mirth Hoyt,

Scott and Erin Shaw, Sergio Sanchez. You have all made my time here so enjoyable, I wish you

all the best. Sergio Sanchez I have always been impressed with your constant thirst for

knowledge and unwavering dedication to the correct answer. You are a truly great person.

This degree would not be possible if it were not for all the facilities workers and staff.

Marie thank you for listening to my issues and all the food you make. Scott and Theresa thanks

for all the conversations about fishing, trucking, X-rays, and life you are missed. Amy and

Page 7: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

vii

Danielle thanks for all the data you collected the past few months. You both have been life

savers and I will not soon forget all the help you gave to me. Jeff Johnstone and Donnie O’Brien

thanks for always listening to my ideas and helping me out with whatever I needed, Jeff I hope

you are enjoying your retirement. Mauro Sardela, Jim Mabon Tim Spila, Mike Marshall,

Changhui Lei, Jian-Guo Wen, Wacek Swiech, Doug Jeffers, and Bharat Sankaran thank you for

all your training in material characterization. Rick Haasch thanks for all your help in collecting

my XPS and AUGER data, conversations with you are always enjoyable and refreshing.

Connie, Beth, Theresa, and Cathy thank you for all the conversations and help with my

paper work. If not for you I am not sure how this department would function. I would like to

thank Connie especially for all her help and guidance. You are always upbeat and positive

which is a welcome change to normal graduate student life. Finally I would like to say thank

you to Eboni. You have been one of the best things in my life for a long time. I look forward to

the next step in our lives together.

Page 8: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

viii

Table of Contents

Chapter 1: Aluminum and Boron Nanoparticles for High-Energy

Applications…………………………………...................................................................... 1

Aluminum Nanoparticles as Energetic Materials……......................................... 1

Synthesis of Aluminum Nanoparticles................................................................... 6

Passivation of Aluminum Nanoparticles……….................................................... 13

Boron Nanoparticles as Energetic Materials......................................................... 29

Synthesis of Boron Nanoparticles........................................................................... 37

Passivation of Boron Nanoparticles....................................................................... 39

Characterization of Boron Nanoparticles.............................................................. 44

References................................................................................................................. 47

CHAPTER 2. Synthesis, Characterization, and Functionalization of Boron

Nanoparticles…………………………................................................................... 55

Introduction.............................................................................................................. 55

Results and Discussion............................................................................................. 57

Experimental Section............................................................................................... 68

References................................................................................................................. 71

Chapter 3: Nanoparticle Composites of Aluminum and Transition Metals. A Scanning

Transmission Electron Micrographic Study of Nanostructured Energetic

Materials………………....................................................................................................... 73

Page 9: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

ix

Introduction.............................................................................................................. 73

Results and Discussion............................................................................................. 74

Experimental Section............................................................................................... 94

References................................................................................................................. 97

Chapter 4. Synthesis of New Aminodiboranate Salts and Their 1,4-Dioxane

Adducts…………………………………………………………………............................. 102

Introduction.............................................................................................................. 102

Results and Discussion............................................................................................. 103

Experimental Section............................................................................................... 108

References................................................................................................................. 117

CHAPTER 5. Preparation and Characterization of Alkaline Earth

Aminodiboranates……………............................................................................... 121

Introduction.............................................................................................................. 121

Results and Discussion............................................................................................. 122

Experimental Section............................................................................................... 132

References................................................................................................................. 144

CHAPTER 6. Structures and Properties of Copper Alkene Complexes. Binding of

Different Isomers of Cyclododecatriene to Copper Triflate................................ 150

Page 10: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

x

Introduction.............................................................................................................. 150

Results and Discussion............................................................................................. 152

Experimental Section............................................................................................... 169

References................................................................................................................. 181

Page 11: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

1

Chapter 1: Aluminum and Boron Nanoparticles for High-Energy Applications

Energetic materials find wide use in propellants, explosives, and pyrotechnics. Although

many energetic materials are organic compounds, the highest energy densities are found in

inorganic materials. Among materials used as fuels, the highest combustion enthalpies of all

known materials are seen for aluminum, boron, beryllium, and magnesium (Table 1.1).1 Some

of these substances (especially aluminum) are already widely used in energetic material

formulations, but all are of interest owing to their high energy densities.

Aluminum Nanoparticles as Energetic Materials. The combustion of aluminum,

which produces liquid Al2O3, is highly exothermic and releases 7.4 kcal per g of Al. Most studies

of the combustion of aluminum are conducted in an oxygen rich environment, but in fuel

mixtures formulations the aluminum is often mixed with a hydrocarbon propellant or a

polymeric binder. Under these circumstances, the aluminum powder primarily reacts with the

organic oxidant to give the oxidation products CO2 and H2O. Aluminum reacts exothermically

with H2O and CO2 to produce H2 and CO, with energy releases of 4.0 and 3.4 kcal per g of Al.2,3

Aluminum particles are a common component of solid propellants. For example, each

solid rocket booster on the space shuttle contains more than 450,000 kg of propellant, which

consists of ammonium perchlorate (oxidizer, 69 % by weight), 20-200 μm aluminum particles

(fuel, 16 %), and polybutadiene acrylic acid acrylonitrile (binder, 12 %), as well as an epoxy

resin (curing agent, 2 %) and iron oxide (catalyst, 0.4 %).4 By adding aluminum particles to the

mixes, the specific impulse generated by the propellant increases by approximately 10 %.5

Specific impulse depends on both the volume of gas produced as well as the heat of combustion

per gram of propellant.6 Interestingly, even the addition of small amounts of aluminum improves

rocket motor operation by reducing combustion instability problems.2,3,7-9

Page 12: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

2

Table 1.1. Heats of combustion of various fuel/oxidant combinations.

Fuel Oxidant(s) Product(s) kcal/mL

Fuel

kcal/g

Fuel

kcal/mol

Fuel

B(s) F2(g) BF3(g) 58.7 25.1 271.3

Be(s) O2(g) BeO(s) 36.3 16.2 145.6

B(s) O2(g) B2O3(s) 32.9 14.1 152.1

Gasoline(l) O2(g) CO2(g), H2O(g) 32.7 11.3 1290.8

B(s) O2(g), F2(g) OBF(g) 31.4 13.4 145.0

Al(s) F2(g) AlF3(s) 28.7 10.6 287.7

Al(s) O2(g) Al2O3(s) 20.0 7.4 200.1

graphite(s) O2(g) CO2(g) 17.2 7.8 94.0

Mg(s) O2(g) MgO(s) 10.3 5.9 143.7

hydrazine O2(g) N2(g), H2O(l) 5.4 5.3 159.3

H2(l) F2(g) HF(g) 4.5 64.8 130.6

H2(l) O2(g) H2O(g) 2.0 28.7 57.8

Page 13: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

3

Although aluminum particles have been used in propellants for several decades,10-15 they

are not a perfect fuel. Aluminum particles tend to burn slowly: the relatively low vapor pressure

of aluminum (vs other more volatile fuels such as hydrocarbons) results in relatively slow gas

phase combustion. The slow combustion rate is further exacerbated by the protective aluminum

oxide shell present on the particles. This shell prevents the aluminum particles from

spontaneously combusting in air, but it also interferes with contact between the oxidizer and the

aluminum during motor operation, resulting in ignition delays. Ignition occurs only after the

particles heat up and the aluminum inside the shell liquefies (m.p. 933 K). The heat-expanded

liquid cracks the aluminum oxide shell (m.p. 2327 K) and the molten aluminum seeps out.

Combustion takes place at a temperature between 2300 and 2700 K, somewhat below the boiling

point of Al (~2750 K) as a detached flame a short distance away from the surface (Figure 1.1).

Under these conditions, most of the aluminum at the burning surface of the propellant

immediately converts to aluminum oxide smoke particles less than 5 μm in diameter, with the

remaining uncombusted aluminum coalescing into larger agglomerates up to 500 μm in diameter.

The agglomerates are carried downstream through the chamber and out the nozzle, where they

remain in a molten state but burn slowly owing to their lower surface to volume ratios. The

presence of these aluminum agglomerates within the chamber volume is beneficial because they

dampen out combustion instabilities due to resonating acoustical waves within the combustion

chamber.

The greatest disadvantage of aluminized propellants is the energy loss associated with

accelerating the large molten agglomerates. When first formed, the aluminum agglomerates are

traveling at relatively low speeds compared to the exhaust gases within the combustion chamber.

When the agglomerates are accelerated to the speed of the gas flow, energy is extracted from the

Page 14: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

4

Figure 1.1. Simplified description of aluminum droplet combustion.16

Page 15: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

5

exhaust gases. In addition, the agglomerates are accelerated not only axially but radially as well,

thus dissipating more energy and causing as much as a 5 % reduction in rocket performance.

Propellants and other energetic materials that use aluminum as the fuel typically employ

aluminum particles with diameters of 5-40 μm, but in recent years there has been increased

interest in aluminum particles in the nanoscale regime. The small sizes of nanoparticles impart

several advantages of use in energetic materials. First, for the same amount of aluminum,

nanoparticles have much larger surface areas. The larger fraction of atoms already on the surface

speeds combustion which, as mentioned above largely takes place on the surface of the particles

rather than the gas phase owing the high boiling point of aluminum (2467 °C). The increase in

surface atoms also results in faster combustion times because of enhanced mass transport rates

between the aluminum and the oxidizer. Second, a smaller particle size results in lower

coordination numbers for the surface atoms and a higher concentration of core defects.17 These

structural features increase the energy of the particle and reduce the activation energy needed for

combustion, which in turn cause faster combustion times.18 Third, the small particle sizes

improve the homogeneity of the propellant mixtures, thus reducing the diffusion distance

between oxidizer and fuel and reducing the ignition delays during combustion.19

Several problems remain to be overcome before aluminum nanoparticles will be useful

constituents in propellant formulations. For example, as seen for micron-sized aluminum

powders, promoting complete combustion and avoiding the formation of metallic agglomerates

remains a challenge. More importantly, passivating the nanoparticles so that they are stable

enough to handle and store for long periods has proven to be very difficult. Some promising

approaches to solving these and other problems will be described in the following sections.

Page 16: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

6

Synthesis of Aluminum Nanoparticles. The techniques for synthesizing aluminum

nanoparticles can be divided into high temperature and low temperature processes. The high

temperature techniques include gas evaporation, plasma chemical synthesis, laser ablation,

electro-explosion, and ion implantation, whereas the low temperature techniques include solution

methods and ball milling.

Synthesis by Gas Evaporation. The most common method to synthesize aluminum

nanoparticles is the evaporation of aluminum from the molten state into a chamber filled with an

inert gas, where the gaseous metal condenses. Inert gas evaporation was developed the 1940s to

make evaporated metal films20,21 but not applied to particles until 1963.22 When aluminum

nanoparticles are synthesized by this technique in a completely oxygen-free environment, they

immediately combust upon exposure to air. If instead the vaporized aluminum is condensed in

an inert atmosphere containing small amounts of oxygen gas, a protective coating of aluminum

oxide forms on the particles. If a sufficient amount of coating is present, the aluminum oxide

layer protects the aluminum core from further oxidation. The resulting aluminum nanoparticles

are thus passivated and can be handled in air without immediate ignition.

The purity of the aluminum starting material, and the type and purity of the inert gas

atmosphere, strongly influence the properties of the aluminum nanoparticles obtained.23,24

Tuning these variables allows control over both the size and the size distribution of the

aluminum nanoparticles. The shape of the particles can vary depending on cooling speed, with

slower cooling rates leading to more spherical nanoparticles. Aluminum particles created by

evaporation and used in propulsion applications are usually nonspherical in shape.5

The morphology and thickness of the aluminum oxide coating on Al nanoparticles

prepared by evaporation have been the subject of several studies.23,25,26 Electron energy-loss

Page 17: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

7

spectroscopy (EELS) showed that 10-40 nm aluminum nanoparticles prepared by inert gas

evaporation had thin (~ 4 nm) amorphous aluminum oxide coatings (Figure 1.2).27 The near-

edge features of the EELS spectra suggested that the particles had a composition of 35%

amorphous Al2O3 and 65% metallic aluminum; these values are in good agreement with those

predicted from the morphology observed by TEM. Extended energy-loss fine structure

spectroscopy revealed that the aluminum atoms in the Al2O3 layer are in tetrahedral

environments.27 In contrast, the aluminum atoms in the oxide layer on an aluminum foil tend to

be in octahedral environments. When the nanoparticles are heated, the aluminum oxide layer

crystallizes into a structure that contains both tetrahedral and octahedral coordinated aluminum

consistent with the formation of the γ-Al2O3 phase. Interestingly, the aluminum atoms in the

metallic core have on average only 9.4 nearest neighbors were determined. This value is far less

than the 12 nearest neighbors that characterize bulk Al, but is in close agreement with the

coordination number deduced from the electron-loss near edge features.

A modified inert gas evaporation method called cryomelting can also be used to make

aluminum nanoparticles.28 In the cryomelting process, the evaporated metal is rapidly condensed

in region cooled to about 70 K. The rapid cooling induces a high rate of nucleation. This

method can produce 20 – 500 nm aluminum nanoparticles in which 60% of the particles are

smaller 70 nm in size, as observed by TEM.28 The amorphous surface layer is 3 nm thick and

EDX showed it to contain only aluminum and oxygen.

Synthesis by Plasma Chemical Synthesis. The plasma chemical synthesis of aluminum

nanoparticles is a single step process that has some advantages over gas evaporation, including

the ability to employ a wide range of aluminum precursors. Pure elements, binary solids, gases,

organic, and inorganic reactants can all be introduced into the plasma to produce nanoparticles.

Page 18: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

8

Figure 1.2. (a) Isolated aluminum nanoparticles coated with 66.5 wt% Al2O3 prepared by inert gas evaporation. (b) Aggregated aluminum nanoparticles with 36 wt% Al2O3.27

Page 19: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

9

Although convenient and flexible, plasma methods often generate particles with broad size

distributions.

Aluminum nanoparticles have been generated an atmospheric pressure plasma torch, in

which micron scale aluminum particles suspended in an aerosol are injected into a microwave

plasma.29 By adjusting the plasma gas flow rate, the aerosol gas flow rate, and the applied

power, it is possible to generate spherical nanoparticles with diameters of 7 - 34 nm.30

Interestingly, many of these spherical aluminum nanoparticles have tails (Figure 1.3). The

authors propose that these tails result when the outside of the molten nanoparticle freezes before

the inside. Subsequent thermal contraction of the outer coat causes molten aluminum to be

ejected through the surface layer, where it subsequently freezes into a tail.

Aluminum nanoparticles generated by a microwave plasma were examined after they had

aged in air for 6 months.31 SEM images showed that the aluminum nanoparticles crystallized as

spheres, and that the Al adopted the usual fcc structure. The spheres themselves were aggregated

into small agglomerates. TEM images showed that the average diameter of the nanoparticles

was 50 nm, but that particles as large as 500 nm were also present. The TEM also showed that

the particles were coated with an amorphous outer layer that was approximately 5 nm thick. Not

surprisingly, XPS studies showed that the surface of the aluminum nanoparticles contained

significant amounts of oxide, but also 10-20% carbon from hydrocarbons. The latter were

proposed to arise during handling.31

Synthesis by Laser Ablation. Pulsed laser ablative deposition (PLD) is an attractive

synthetic method owing to its ability to produce nanoparticles with a narrow size distribution and

a low level of impurities.32-34 Aluminum nanoparticles with diameters of tens to 500 nm of

various shapes can be prepared by irradiating an aluminum foil with 50 fs pulses of a 0.8 μm

Page 20: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

10

Figure 1.3. Aluminum nanoparticles prepared by plasma torch method showing spheres and tails.30

Page 21: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

11

wavelength laser beam.35 For any particular aluminum source and substrate, the size and

number of the nanoparticles generated depends only on the laser power per unit area. This

dependence can be severe. For example, no nanoparticle formation is observed at laser

irradiances below 3 x 1012 W/cm3 (at which there is insufficient energy to form a nanoparticle-

generating plasma) and above 5 x 1014 W/cm3 (at which the initially formed plasma is too high in

temperature to cool and aggregate before expansion disperses the plasma constituents). Between

these limits, the number of nanoparticles produced per unit area of substrate first increases as

irradiance increases and then tails off, with the greatest yield being observed at a irradiance of

5.3 x 1013 W/cm2.35

Synthesis by Electro-explosion. Electro-explosion of metal wires has only recently been

seriously applied to make aluminum nanoparticles. In electro-explosion, a brief but powerful

current pulse creates an electromagnetic field around the wire that holds it together while it is

superheated to tens of thousands of degrees. When the current ceases, the electromagnetic field

disappears and the wire fragments into nanosized particles. The shapes and sizes of the resulting

particles depend on many factors, such as the shape and size of the wire, the voltage, and the

nature of the electrical pulse. Typically, aluminum nanoparticles synthesized by electro-

explosion are about 50 nm in size, irregularly shaped, and 92 – 95% aluminum, with the balance

being oxygen if the explosions are carried out in air.36

The aluminum nanoparticles synthesized by electro-explosion do not oxidize completely

when they are ignited on a ceramic sheet, possibly because the rate of self-sintering of the

particles is faster than the burn rate. The aluminum nanoparticles react with water at 50 °C to

produce aluminum oxide and hydrogen gas. When aluminum nanoparticles made by electro-

explosion are mixed with an oxidizer such as ammonium perchlorate or potassium nitrate, the

Page 22: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

12

burn rate at low concentrations of aluminum nanoparticles was nearly equal to that of industrial

aluminum powders. At higher concentrations, however, the burn rate increased up to 20 times,

reaching a maximum at 60 wt. % aluminum nanoparticles. Interestingly, for industrial aluminum

powder, mixtures richer than 40% aluminum were not combustible.36

Phung et. al. studied the surfaces of aluminum nanoparticles made by electro-explosion

in the 20 – 260 nm size regime, with an average size of 100 nm.37 They report that the

nanoparticles are spherical when viewed by TEM, but Ivanov and Tepper noted that the beam

from a TEM causes irregularly shaped aluminum nanoparticles made by this method to form

spheres due to melting and sintering within several seconds if the beam is not of sufficiently low

power.36 Using HRTEM, Phung et. al. observed lattice fringes that indicated that the aluminum

core was crystalline; in contrast, the upper 3 nm of the surface was amorphous. These results

are consistent with those obtained from aluminum particles made by other methods.28,31,38 The

XEDS spectrum showed that the oxygen concentration was high in the surface layer, but dropped

off suddenly at the surface-core interface. This result is consistent with the presence of an Al2O3

surface layer that is not miscible with the aluminum core.

Synthesis in Solution. Treatment of aluminum chloride with lithium aluminum hydride in

mesitylene at 164 °C affords aluminum nanoparticles according to the equation:39

3 LiAlH4 + AlCl3 ―→ 4 Al + 3 LiCl + 6 H2

The nanoparticle aggregates made by this method were 110 – 210 nm in diameter. As isolated,

the aluminum nanoparticles are mixed with lithium chloride, but most the lithium chloride can be

Page 23: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

13

removed by washing. This method proved to be inconvenient for scale up and, even after being

washed, the nanoparticles still contained measurable levels of carbon, oxygen, and chlorine.

Purer aluminum nanoparticles can be obtained on a larger scale by decomposing isolated

samples of the alane amine adduct H3Al(NMe2Et) in mesitylene at164 °C:39

2 H3Al(NMe2Et) ―→ 2 Al + 3 H2 + 2 NMe2Et

The aluminum nanoparticles aggregates synthesized by this method proved to be smaller as well,

with diameters of 44 – 82 nm. When 1 mol % of Ti(iOPr)4 was added to the reaction medium,

the decomposition of the alane amine occurred at temperatures as low as 100 °C, and the average

aggregate size was 50 – 75 nm. Compacting the aluminum nanoparticles in a press at 350 MPa

under argon caused the grains to increase in size; after three months, grains larger than 200 nm

were obtained even at room temperature. Grain growth under these conditions does not occur

when aluminum nanoparticles are prepared by other methods, probably because the latter have a

larger concentration of oxide impurities that stabilize the grain-boundaries.25,26

Aluminum nanoparticles in the 65 – 500 nm size regime have also been obtained

thermolysis of H3Al·NMe3 in refluxing toluene or xylene in the presence of TiCl4.40 Increasing

the amount of TiCl4 results in the formation of smaller aluminum nanoparticles, probably

because the TiCl4 increases in the number of initiation sites. For example, addition of 0.16 %

TiCl4 produced 500 nm particles, whereas addition of 20 % TiCl4 yielded 150 nm particles.

Passivation of Aluminum Nanoparticles. The small sizes of aluminum nanoparticles

make them particularly susceptible to excessive oxidation while being stored prior to use.

Typically, the thickness of an oxide coating on an aluminum particle ranges from 1.7 to 6.0 nm,

Page 24: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

14

irrespective of the size of the particle.41 For sufficiently small particles, this oxide layer will

constitute the majority of the mass. As a result, much attention has been devoted to modifying

the aluminum nanoparticles in order passivate the surface against the formation of an oxide

overlayer, and thereby obtain longer shelf lives and better burn properties. If the passivating

coatings also have an affinity for the binder material, then mixing problems can be resolved as

well. These coatings include but are not limited to small molecules, polymers, and transition

metals.

Passivation of Aluminum by Small Molecules. Dinitrogen, carbon dioxide, steric acid,

oleic acid, and nitrocellulose have been examined as possible passivating agents for aluminum

nanoparticles made by the electrical explosion of wires.42-44 Aluminum nanoparticles treated

with dinitrogen and carbon dioxide were not effectively passivated, and oxidized when exposed

to air. The nitrocellulose proved to be a powerful oxidizer of aluminum and the resulting

nanoparticles did not perform well in aging studies. Neither steric acid, oleic acid, nor

nitrocellulose coated the nanoparticles particularly well: many of the nanoparticles showed no

organic coating as judged by TEM. A few particles, however, exhibited an organic layer on top

of the aluminum oxide coating (Figure 1.4).44 The physical properties of the coated

nanoparticles prepared by wire explosion are summarized in Tables 1.2 and 1.3.

Passivation of Aluminum by Fluorinated Molecules. Nanoparticles coated with

organofluorine passivating agents could prove useful in propulsion applications because

aluminum fluoride has a very exothermic heat of formation. Specifically, the conversion of

metallic aluminum to Al2O3 releases 7.4 kcal/g of Al whereas the conversion of aluminum to

AlF3 releases 13.3 kcal/g of Al.

Page 25: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

15

Figure 1.4. Aluminum nanoparticles prepared by electrical explosion of wires and passivated with an organic coating.42

Page 26: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

16

Table 1.3. Combustion properties of coated aluminum nanoparticles prepared by wire

explosion.44

Table 1.2. Surface areas and percent metal content of fresh and aged organic-coated aluminum nanoparticles prepared by wire explosion.44

Page 27: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

17

In a modification of Higa’s method,40 it has been claimed that passivated aluminum

nanoparticles can be obtained by adding Ti(iOPr)4 to a solution an alane-amine such as

AlH3·NMe3 or AlH3·(N-methylpyrrolidine) at room temperature, stirring for 5 min to 2 h, and

adding a fluorinated carboxylic acid. Notably, in this procedure the decomposition of the alane

was carried out for short times at room temperature, whereas Higa’s method calls for heating the

alane and the titanium catalyst to between 82 and 141 °C for 25 minutes. The evidence suggests

that a major product of this room temperature procedure is an aluminum(III) carboxylate, along

with elemental aluminum. Significantly, a minimum carboxylic acid to aluminum mole ratio of

4.9:1 was required to make air stable nanoparticles.45 This finding is consistent with the view

that, even after the alane is stirred with the titanium catalyst for 2 h at room temperature (and

certainly after only 5 min), much of the AlH3⋅(amine) was still present when the carboxylic acid

was added. The amount of aluminum in the resulting nanoparticles was determined by base

hydrolysis46 to be only 15 % Al by weight.45 This value is well below the percentage of active

aluminum in common aluminum nanoparticles, and also suggests that considerable Al3+ (i.e.,

non-metallic) aluminum is present.

In a follow-up study, the catalyst for the decomposition of H3Al⋅(N-methylpyrrolidine)

was changed from Ti(iOPr)4 to TiCl4, and the stirring time was increased to 4 h before addition

of the perfluorocarboxylici acid. Under these conditions, less carboxylic acid (1 for every 13 Al

atoms) is needed to passivate the nanoparticles.47 Jouet et al. hypothesize that less carboxylic

acid is needed because the aluminum nanoparticles are larger and therefore have smaller surface

areas, but more likely is that the longer stirring time increases the conversion yield of alane to

aluminum, which is a slow reaction at room temperature. The resulting particles have a

maximum of 33 % Al by weight, which is still low and suggests that aluminum(III) carboxylates

Page 28: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

18

may be present in these samples as well. SEM images show that the particle diameters are 20-

200 nm (Figure 1.5).47 The presence of the perfluoroalkyl carboxylic acid in the final material

was confirmed by IR spectroscopy (Figure 1.6). Even though the IR spectrum was obtained in

air, there was no evidence of the presence of aluminum oxide or chemisorbed oxygen, in the

nanoparticles, as judged from the absence of IR stretches at ~610 cm-1 and ~825-865 cm-1

indicative of oxide formation, and the vibrational mode at ~590 cm-1 characteristic of

chemisorbed oxygen.48

Small scale shock reactivity tests show that the Al nanoparticles coated with C13F27CO2

perform nearly as well as Valimet H5 (Valimet H5 is an aluminum powder with particles sizes ~

7 μm) (Figure 1.7). 47 This result is significant because the Valimet H5 Al particles are 99.7 wt

% aluminum whereas the Al-C13F27OO particles are only 33 wt % aluminum.

Passivation of Aluminum by Polymers. Aluminum nanoparticles embedded in a poly-

paraxylylene matrix can be prepared co-condensation of aluminum and paraxylylene biradicals

onto a cold (-196 °C) surface; the biradicals are generated by pyrolysis of paracyclophane. 19

When the substrate is warmed, the paraxylylene biradicals polymerize and trap the aluminum

nanoparticles inside the polymer matrix. Such matrices are known to be porous, so that all the

nanoparticles should be fully accessible to air.

Changes in the electrical resistivity of the aluminum-polymer composites upon exposure

to air were interpreted to mean that aluminum-polymer composites with 12 wt. % aluminum

oxidize very little in air, whereas aluminum-polymer composites with 8 wt. % aluminum oxidize

completely.49 This somewhat surprising result was attributed to interparticle charging (i.e,

electron exchange between ensembles) that occurs only when the particle concentrations are high

and the interparticle distances are short. The resulting electrical fields are proposed to prevent

Page 29: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

19

Figure 1.5. TEM of nanoparticles prepared by decomposition of H3Al·(N-methylpyrrolidine) and subsequent treatment with C13F27OOH.47

Page 30: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

20

Figure 1.6. Bottom: IR Spectrum of C12F27OOH Top: IR spectrum of aluminum nanoparticles passivated with C13F27CO2H showing carbonyl C=O stretch at 1754 cm-1 and C-F stretches at 1200 and 1142 cm-1. The absence of bands at 3076 and 2930 cm-1 suggest that the C13F27CO2H acid has been converted to its carboxylate conjugate base.47

Page 31: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

21

Figure 1.7. Shock reactivity tests of C13F27OOH passivated aluminum nanoparticles.47

Page 32: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

22

oxygen ion movement towards the metal particles. Such speculations are unwarranted in the

absence of better evidence of the different oxidation resistance of the different composite

compositions.50.

Aluminum nanoparticles can also be coated with polymers by Ziegler-Natta methods.

The Al nanoparticles are first treated with the TiCl4 catalyst, which react with surface hydroxyl

groups, and then are coated by addition of ethylene and the co-catalyst triethylaluminum (Figure

1.8). Nanoparticles coated by this method agglomerate to 2 μm diameter aggregates, as

measured by light scattering.51 Polymers coated by this method are effectively passivated, as

shown by accelerated aging tests (Figure 1.9).

Aluminum nanoparticles can also be coated with thermoset polymers such as Bisphenol

A, hydroxyl-terminated polybutadiene, and Zonyl (a fluoropolymer). In this technique, the

hydroxyl groups on the nanoparticles are first treated with an organic di-isocyanate and then with

a hydroxylated polymer. This method is effective for coating micron sized particles, but does

not coat nanoparticles well.52

Aluminum nanoparticles generated by plasma methods can also be coated with

poly(methyl methacrylate) by injecting methyl methacrylate monomer into the reactor slightly

downstream of the plasma region where the nanoparticles are formed. The resulting coating is

~100 nm thick; some unpolymerized monomer is present in the coatings.52

Passivation of Aluminum by Other Elements. Aluminum nanoparticles generated by laser

ablation or a DC plasma can be coated with carbon by introducing ethylene to the argon quench

flow.53 The resulting nanoparticles have an average mobility diameter of 80 nm. TEM images

clearly show a thin ~5 nm coating (Figure 1.10). Single particle mass spectroscopy showed that

the nanoparticles are largely aluminum but also contain small amounts of carbon

Page 33: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

23

Figure 1.8. Method of coating aluminum nanoparticles with polyethylene.52

Page 34: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

24

Figure 1.9. Accelerated aging tests of aluminum particles before and after being coated with polyethylene.52

Page 35: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

25

Figure 1.10. TEM images of aluminum nanoparticles coated with carbon in a DC plasma.53

Page 36: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

26

and hydrogen, but no oxygen.54 When exposed to heated air and sampled by single particle mass

spectroscopy, the nanoparticles are resistant to oxidation up to 700 °C (Figure 1.11).

Transition metals have also been employed as coatings to passivate aluminum

nanoparticles. Coatings of palladium, 55 silver,55 and nickel55-58 were prepared by treating bare

aluminum nanoparticles with the corresponding metal acetylacetonate, whereas coatings of gold

were obtained from gold(I) chloride dimethyl sulfide.55 The metal salts are reduced by the

aluminum nanoparticles, thus depositing metal onto the surface. The chemical and physical

properties of these metal-coated nanoparticles are summarized in Table 1.4. The active

aluminum content is highest for the nickel coated nanoparticles. Particles coated with palladium,

silver, and gold, have lower active aluminum contents, probably because of the higher molar

masses of these elements. The nickel passivated aluminum nanoparticles are also more resistant

to air oxidation than air-passivated aluminum nanoparticles; in other words, the Al2O3 layer that

forms when the aluminum is exposed to air is more porous to oxygen than the nickel coating.

Nickel containing aluminum nanoparticles have also been made by electro-explosion of

aluminum-nickel composite wires.42,44 When aluminum nanoparticles are made in this way, the

nickel does not form an outer coating but instead separates into a small oxidized region of the

particle, as observed by EDX. Nickel does not increase the air stability of aluminum

nanoparticles made in this manner.

Aluminum diboride has also been investigated as an alternative to aluminum oxide as a

passivating layer.42,44,59 Aluminum diboride is interesting as a passivating layer because, unlike

the inert aluminum oxide, it combusts exothermically with an energy release of 9.9 kcal/g. It has

been claimed that aluminum nanoparticles coated with AlB2 can be made by electro-explosion of

wires coated with boron. Although electron probe microanalysis showed that AlB2 was present

Page 37: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

27

Figure 1.11. Single particle mass spectroscopy of carbon-coated aluminum nanoparticles after heating in air.53

Page 38: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

28

Table 1.4. Chemical and physical properties of aluminum nanoparticles passivated with metals.55

Page 39: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

29

on the surface of the particles, the formation of a coating was not proved. Although the Al/AlB2

particles are as susceptible to air oxidation as Al particles made by electro-explosion and then

passivated with Al2O3, upon combustion they produce more energy than their Al2O3 coated

counterparts.

Passivation of Aluminum by Ball Mill Blending. Mechanical ball milling has been used

to blend aluminum with magnesium60,61 and carbon62-65 in order to alter its chemical properties

and combustion behavior. The studies to make blends with magnesium used particles tens of

micrometers in size. Intermetallic Al-Mg phases start to form at magnesium concentrations

above 30%, but beneficial effects are seen even below this threshold: the ignition temperature

for a mixture containing < 30 % Mg is some 1300 K below that of pure aluminum. The

combustion products consist exclusively of oxide species, indicating that complete combustion

occurs; in contrast, combustion of otherwise similar particles of pure aluminum was incomplete

and aluminum metal could be detected among the products.

Interestingly, aluminum nanoparticles ball milled with carbon were much more reactive

than standard aluminum nanoparticles, and had to be handled under hexane to keep them from

spontaneously combusting in air.63

Boron Nanoparticles as Energetic Materials. Boron, the lighter congener of

aluminum, is a highly attractive high-energy material. As a fuel additive, boron boasts the

second greatest heating value of any element.66 Another attractive feature of boron is that it and

its oxidation product B2O3 are non-toxic. The melting point of boron, 2450 K, is far too high for

it to be used as a liquid fuel, but it has potential as a liquid dispersion or solid composite.67

Despite the desirable properties of boron, the incorporation of boron nanoparticles into

propellants is not widely practiced, for reasons we will discuss below. Although molecular

Page 40: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

30

boron chemistry is well studied, routes to high energy nanoparticles are few. Below, we

summarize the efforts that have been made to improve the combustion efficacy of boron-

containing materials and to synthesize boron nanoparticles.

The combustion of boron to B2O3 releases 14 kcal/g, which is considerably more than the

7.2 kcal/g released by aluminum and the 10.6 kcal/g by n-octane.68 Compared to JP4, a widely

used conventional liquid hydrocarbon fuel, the energy content of boron is 35% greater

gravimetrically and 300% greater volumetrically. Although from a thermodynamic perspective,

boron is an attractive component of propellant, it suffers from several kinetic problems: solid

boron samples bear B2O3 coatings that limit diffusion of oxygen, the combustion mechanism

involves a long-lived metastable intermediate (HOBO) that reduces the energy released by the

combustion reaction, and B2O3 condenses from the gaseous to the liquid state too late in the

combustion process (i.e., after it leaves the thrust chamber), thus robbing the motor of some of its

propulsive energy. One of the reasons for the current interest in boron nanoparticles is that they

should suffer less from the first of these issues, although finding solutions to the other kinetic

problems will be more challenging.

The ignition and combustion of boron and aluminum follow different mechanistic

pathways owing to the different melting and boiling points of the reactants and products. As

described in section 1.0, during the combustion of aluminum particles, the solid aluminum oxide

shell is broken apart by the thermal expansion of the liquid aluminum core.69 As the liquid

aluminum seeps through the cracks in the oxide shell and vaporizes, it reacts with the oxidizer in

a homogeneous gaseous envelope around the particle. The temperature becomes high enough to

melt the Al2O3, but the energy consumed during melting is re-released as the particles travel

downstream from the combustion zone and cool.

Page 41: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

31

In contrast, the melting point of B2O3 is lower than that of boron, so that the oxidizer can

access the inner boron particle for ignition, but further oxidation is impeded by the oxide

generated, which must evaporate to permit additional oxidant to reach the unreacted boron

(Figure 1.12).66,70 This oxide formation/evaporation cycle continues until the energy released

from the oxidation exceeds that absorbed by evaporation and heat loss. At this point, the

remaining oxide evaporates and the boron particle – now much smaller and less energetic –

reignites and combusts heterogeneously on the surface of the particle. The high melting point of

boron means that, unlike aluminum, the nanoparticle itself never liquefies. Thus, the combustion

of boron particles is a two-stage process: in the first stage the oxide layer controls the ignition

process, and in the second state the oxide layer evaporates quickly enough to expose fresh boron

surface and sustain the oxidation reaction.70

The burning times of crystalline boron at ambient pressures can be decreased by

increasing the temperature, increasing the amount of oxygen, and increasing the amount of water

vapor. Boron particles smaller than 20 μm in diameter have sufficiently short lifetimes to be

suitable for high-speed air-breathing propulsion applications.71 Water vapor decreases the

ignition delay by assisting in oxide layer removal.

Ignition of boron occurs at 1500 + 70 °C and is independent of gas composition, heating

rate, and preliminary annealing in air. Combustion of an electrically-heated crystalline boron

filament has been proposed to occur in two distinct stages.72 The first stage of combustion

coincides with the phase transition from α- to β-rhombohedral boron (around 1500 ºC). In this

stage, oxygen gas dissolves in the boron filament until a supersaturated solution is formed.

When the oxygen content approaches supersaturation, the oxygen dissolution rate decreases,

which causes a cooling effect. The second stage begins when the supersaturated solution

Page 42: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

32

Figure 1.12. Ignition model illustrating the chemical species and processes involved during heating of a boron particle.66

Page 43: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

33

undergoes a phase separation in which the oxygen gas is released, forming spherical voids within

the boron (Figure 1.13). At this point, combustion of the boron filament is reinitiated.

The chemistry occurring at the particle surface under both fuel-rich and oxygen-rich

conditions has been modeled by assuming that the initial reactant concentrations are equal to

calculated equilibrium concentrations at the boundary layer.73 Studies of the homogeneous gas

phase combustion chemistry afforded rate constants for the relevant reactions, which showed that

the dominant species in the boundary layer are HBO, B2O2, and BO under fuel-rich conditions,

and HOBO, B2O3, and BO2 under oxygen-rich conditions. In addition, hydrogen-containing

species such as H2O, OH, H, and H2, which are present at temperatures above 1800 K, initially

accelerate the combustion process, but when their concentrations increase further, they induce

the formation of HOBO, a thermodynamic sink that is favored over B2O3 at all but the highest

combustion temperatures (T > 3000 K). Yetter has summarized the critical reactions that are

relevant to the combustion (Figure 1.14).74

The model for the high-temperature gas phase combustion of boron has been further

refined by the addition of other important reactions such as the isomerization of HBO to BOH.67

Both of these species appear to be important to the gas phase combustion process. The species

HOBO and HBO have been identified as the major products at times later than 10-5 s after

commencement of combustion, confirming that these species are thermodynamic sinks, in

agreement with earlier findings. These modeling studies considered only gas-phase combustion

reactions; but HOBO and HBO may also be important energy traps for heterogeneous

reactions.67

The shapes of the boron nanoparticles can affect the ignition process: for example,

nonspherical symmetry may lead to local thinning of the liquid B2O3 film. Surface tension

Page 44: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

34

Figure 1.13. SEM image of the cross-section of a boron filament quenched during the first-stage combustion in air, illustrating the formation of spherical voids upon dissolution of oxygen from boron particles.72

Page 45: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

35

Figure 1.14. Key reactions and intermediates as judged from modeling studies of (a) B/H/O/C combustion processes and (b) in B/H/O/C/F combustion processes.74

Page 46: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

36

gradients and repulsive van der Waals forces can lead to rupture of the B2O3 film and subsequent

ignition.75,76 The surface tension of B2O3 increases with temperature, inducing a net flow of

material toward the hotter parts of the particle surface. Tangential stresses induced by the

surface tension create instabilities at the interfaces of the B2O3 coating. For example, the surface

of the liquid B2O3 coating can become wavy, although by itself this effect is not sufficient to

rupture the coating.

The energy released when boron is used as a fuel is significantly increased if the oxidizer

is a molecule that contains fluorine, such as those based difluoroamines. Using organo-fluorine

oxidants circumvents the formation of boron oxides and oxyhydrides such as HOBO that prevent

the full release of the combustion energy of boron. Kinetic models have shown that the

oxidation of boron particles in the presence of O and F proceeds via the gas-phase formation of

OBF(g) instead of HBO/BOH.74 Both experimental data and kinetic models show that amount

of OBF(g) formed is sensitive to the fluorine/oxygen ratio.77 Also, the reaction rates and heat

release rates both increase in fluorine environments. Ab initio methods show that metastable

intermediates HBO2, HBO, and B2O2 can react with HF to form FBO, although the reaction rates

are probably slower than those used in the modeling studies.78,79 Computational methods have

recently been used to augment experimental data to determine the energetics of boron

combustion reactions.80

The combustion of boron-containing alloys made from ball-milling has also been studied.

Both B-Ti and B-Mg alloys are highly reactive, although the reaction pathways differ from those

of pure boron nanoparticles. The principal products of the combustion of B-Ti alloys are boron

oxides, titanium oxides, and TiB2.81

Page 47: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

37

Synthesis of Boron Nanoparticles. There are many studies of the synthesis and

properties of nanoparticles of boron-containing materials such as boron carbides, boron nitrides,

and transition metal borides, but there are remarkably few studies of boron nanoparticles. Bulk

samples of boron can be obtained by the reduction of boron oxides with alkali or alkaline earth

metals, by the reduction of boron halides with alkali metals or with hydrogen, or by the

electrolytic reduction of boron compounds.82 These methods often result in the formation of

boron colloids or powders, which can be purified and converted into bulk samples by high

temperature methods such as zone melting. Several of these techniques have been modified to

afford boron nanoparticles.

Bulk samples of impure boron were first made in the 1800s by chemical reduction of

B2O3 or boric acid with metals such as potassium. In a similar way, boron-containing colloids

were claimed to result from reduction of B2O3 with Mg, Na, or K, followed by treatment with

HCl and washing with distilled water.83 Later workers had difficulty reproducing this synthesis,

but showed that boron colloids could be obtained by heating a 3:1 mixture of B2O3 and Mg to

fusion in an iron vessel, after which the solid product was ground, refluxed with boiling water,

and treated with concentrated HCl.84 The presence of impurities such as sodium and iron salts

often leads to coagulation of the colloids, which has continued to be a problem in recent

syntheses.

The reduction of boron halides has long been employed to prepare elemental boron.

Treatment of BCl3 with hydrogen in an electric arc85 or on hot tungsten or tantalum substrate86-88

affords boron in high yields.89 Heating a mixture of BCl3 and H2 with a CO2 laser yielded non-

agglomerated, low internal porosity nanoparticles on the order of 50-200 nm in size and with

moderately spherical shapes (Figure 1.15).90

Page 48: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

38

Figure 1.15. Bright field TEM images of non-agglomerated 50-200 nm boron nanoparticles formed by laser heating of BCl3 and H2, showing (a) typical nanoparticles with irregular shapes and (b) atypical nanoparticles with spherical shapes.90

Page 49: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

39

The reduction of BCl3 with Na at elevated temperatures in the gas phase affords boron

nanoparticles that are encapsulated in NaCl.91 The BCl3 is heated in a diffusion flame reactor in

which one counterpropagating flow contains Na in Ar, and the other contains BCl3 in Ar. The

NaCl shell prevents agglomeration of the boron nanoparticles and protects them from air

oxidation.

Boron nanoparticles can also be prepared by thermolysis of diborane, B2H6. The reaction

to generate boron and H2 is exothermic; in contrast, the thermolysis of boron halides to boron

and halogen is endothermic. In an early study, high purity boron nanoparticles were prepared by

heating diborane in a quartz tube furnace to 700 °C, using helium as a carrier gas.92 The

nanoparticle diameters as ascertained by electron microscopy ranged from 25 nm to 500 nm.

When diborane is thermolyzed at higher temperature, 1200-1500 °C, using a CO2 laser, spherical

nanoparticles 30 to 40 nm in diameter are obtained (Figure 1.16).90 These particles, which are

~96.5% boron by mass, agglomerate into chains and clusters, but the agglomeration processes

can be avoided by conducting the thermolysis at temperatures above the 2250 °C melting point

of boron. Passing diborane through a 20 amp electric arc for 10 seconds also induces thermal

decomposition to give boron nanoparticles, which have diameters of 55-95 nm and uniform

shapes.93

Boron nanoparticles can be prepared in liquid media at ambient temperatures by pulsed-

laser vaporization. The higher pressures and increased collision rates between vaporized atoms

and liquid promote the nucleation of particles.94 Belt-shaped nanoparticles have also been

fabricated by pulsed laser deposition (PLD) at low vacuum.95

Passivation of Boron Nanoparticles. Surface passivation of the boron nanoparticles can

be achieved by treating them with TiCl4 and triethylaluminum, followed by addition of ethylene

Page 50: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

40

Figure 1.16. Bright field TEM images of as-synthesized boron nanoparticles synthesized by laser heating of B2H6.90

Page 51: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

41

to generate a polyethylene coating by a Ziegler-Natta polymerization reaction.96 More recently,

it has been reported that surface-passivated boron nanoparticles can be obtained by sodium

naphthalenide reduction of BBr3 in 1,2-dimethoxyethane, followed by treatment with n-

octanol.97 The initial reduction step is proposed to afford boron nanoparticles with surface

bromine groups, which react with the alcohol to give the final product, which is a yellow oil.

The resulting particles range in size from 1 to 45 nm, although most have diameters between 1

and 3 nm (Figure 1.17). There is, however, no evidence that these particles are elemental boron,

and in fact subsequent studies suggest that the yellow oil consists largely of the molecular

species tri(octyloxy)borane.

There are a few studies of the synthesis of boron-rich nanoparticles that also contain other

elements. For example, the gas phase pyrolysis of carboranes such as C2B4H6 and C2B10H12

generates solids of stoichiometry (C2B4H2)n and (C2B10H4)n, respectively, as determined by

elemental analysis.98 A kinetic model was developed for the growth of the particles as a function

of temperature, carborane pressure, and reaction time. This model was consistent with the

formation of nanoparticles that were 10-30 nm in size.

The ignition and combustion problems characteristic of pure boron nanoparticles have

prompted investigation to determine whether nanocomposites containing boron and other

elements exhibit more favorable combustion kinetics. Nanocomposites of B-Ti and B-Zr, for

example, can be synthesized by ball-milling powdered mixtures of the pure elements and

interrupting the milling process just prior to the triggering of a spontaneous, exothermic reaction

(Figure 1.18). This technique, known as arrested reactive milling, leaves the nanocomposite

materials in a highly reactive state that can be passivated and size-controlled by addition of a

Page 52: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

42

Figure 1.17. Bright field TEM images of (a) 2-4 nm nanoparticles and (b) 18-20 nm particles synthesized by reduction of BBr3 with sodium naphthalide followed by alcoholosis with octanol.97

Page 53: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

43

Figure 1.18. A plot of milling time versus temperature showing the stopping point for arrested reactive milling, at which the solid state reactants are left in a high potential energy state.99

Page 54: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

44

liquid medium.99 The combustion of these nanocomposites is complete and rapid, possibly

because a different combustion pathway is followed.

Characterization of Boron Nanoparticles. Combustion tests of boron nanoparticles

synthesized in various ways have been carried out in a lab-scale propulsion system.100 The

results were correlated with physical properties such as average particle diameter, average

surface area, active content, and oxide thickness, as determined by SEM, TGA, DSC, and BET.

The study revealed that it is also important to take into account other properties, such as particle

size distribution, degree of agglomeration, reactivity, and thermal effects. One finding is that

boron particles passified with fluorinated coatings perform better in combustion studies than

uncoated particles.

Near-IR and visible absorption spectroscopy have been used to characterize boron

nanoparticles embedded in dielectric oxidizing media.101 Boron particles from various sources

were first sieved to narrow the size range, then dispersed into nitrocellulose or Teflon thin films

in the presence of surfactants to prevent agglomeration (Figure 1.19). Because the nanoparticles

are more strongly absorbing than the bulk material, single-particle absorption cross sections (σ)

can be calculated from Beer’s law. The method is sufficiently sensitive that nanoparticle

loadings of 3%, 1%, and 0.5% can be distinguished (Figure 1.20). Particle sizes could also be

distinguished because, at the same weight percent loading, smaller particles tended to absorb

more light than their larger counterparts, although exceptions to this rule were noted.

Page 55: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

45

Figure 1.19. Photomicrographs of boron/Teflon samples with and without surfactant. Larger aggregates are visible in the absence of surfactant, as seem from the image on the left.101

Page 56: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

46

Figure 1.20. Absorption spectra of B/Teflon samples at 3%, 1%, and 0.5% loadings for 4 different commercial source of boron particles.101

Page 57: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

47

References

1. Handbook of Chemistry and Physics; 77th ed.; Lide, D. R., Ed.; CRC Press: New York, 1996.

2. Brooks, K. P.; Beckstead, M. W. J. Propul. Power 1995, 11, 769-780.

3. Olsen, S. E.; Beckstead, M. W. J. Propul. Power 1996, 12, 662-671.

4. Fundamentals of Solid-Propellant Combustion; Kuo, K.; Summerfield, M., Eds.; AIAA: New York, 1984; Vol. 90.

5. Price, E. W.; Sigman, R. K. CPIA Publication 1999, 691, 227-248.

6. Kubota, N. Propellants and Explosives 2nd ed.; Wiley-VCH: Weinheim, Germany, 2007.

7. Bucher, P.; Yetter, R. A.; Dryer, F. L.; Parr, T. P.; Hanson-Par, D. M.; Vicenzi, E. Twenty-Sixth Symposium (International) on Combustion, Pittsburgh, PA, 1996; p 1899-1908.

8. Povinelli, L. A.; Rosenstein, R. A. AIAA Journal 1964, 2, 1754-1760.

9. Price, E. W. AIAA Journal 1971, 9, 987-990.

10. Kraeutle, K. J., CPIA Pub., 1976; p 461-477.

11. Kraeutle, K. J., CPIA Pub., 1972; p 325-340.

12. Price, E. W. Prog. Astronaut. Aeronaut. 1984, 90, 479-513.

13. Price, E. W. AGARD Conference Proceedings 1979, 259, 14-1-14-15.

14. Price, E. W.; Sigman, R. K., CPIA Pub., 1977; p 195-208.

15. Crump, J. E.; Prentice, J. L.; Kraeutle, K. J. Combust. Sci. Technol. 1969, 1, 205-23.

Page 58: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

48

16. Melcher, J. C.; Burton, R. L.; Krier, H. CPIA Publication 1999, 691, 249-258.

17. Schmid, G. Clusters and Colloids From Theory to Applications; VCH Publishers, Inc.: New York, 1994.

18. Aumann, C. E.; Skofronick, G. L.; Martin, J. A. J. Vac. Sci.Technol. B 1995, 13, 1178-1183.

19. Pivkina, A.; Ulyanova, P.; Frolov, Y.; Zavyalov, S.; Schoonman, J. Propellants, Explos., Pyrotech. 2004, 29, 39-48.

20. Beeck, O. Rev. Mod. Phys 1945, 17, 61-71.

21. Beeck, O.; Smith, A. E.; Wheeler, A. Proc. Roy. Soc. (London) 1940, A177, 62-90.

22. Kimoto, K.; Kamiya, Y.; Nouoyama, M.; Uyeda, R. Jpn. J. Appl. Phys., 1963, 2, 702-13.

23. Sun, X. K.; Xu, J.; Chen, W. X.; Wei, W. D. Nanostruct. Mater. 1994, 4, 337-44.

24. Puszynski, J. A. Proc. Int. Pyrotech. 2002, 29th, 191-202.

25. Eckert, J.; Holzer, J. C.; Ahn, C. C.; Fu, Z.; Johnson, W. L. Nanostruct. Mater. 1993, 2, 407-13.

26. Sanchez-Lopez, J. C.; Fernandez, A.; Conde, C. F.; Conde, A.; Morant, C.; Sanz, J. M. Nano. Mat. 1996, 7, 813-822.

27. Fernandez, A.; Sanchez-Lopez, J. C.; Caballero, A.; Martin, J. M.; Vacher, B.; Ponsonnet, L. J. Microscopy 1998, 191, 212-220.

28. Champion, Y.; Bigot, J. Nanostruct. Mater. 1998, 10, 1097-1110.

29. Phillips, J.; Perry, W. L.; Kroenke, W. J. US 2001-17289, 2004.

Page 59: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

49

30. Weigle, J. C.; Luhrs, C. C.; Chen, C. K.; Perry, W. L.; Mang, J. T.; Nemer, M. B.; Lopez, G. P.; Phillips, J. J. Phys. Chem. B 2004, 108, 18601-18607.

31. Luo, P.; Nieh, T. G.; Schwartz, A. J.; Lenk, T. J. Mater. Sci. Eng. 1995, A204, 59-64.

32. Pronko, P. P.; Zhang, Z.; VanRompay, P. A. App. Surf. Sci. 2003, 208-209, 492-501.

33. Teghil, R.; D'Alessio, L.; Santagata, A.; Zaccagnino, M.; Ferro, D.; Sordelet, D. J. App. Surf. Sci. 2003, 210, 307-317.

34. Ullmann, M.; Friedlander, S. K.; Schmidt-Ott, A. J. Nanopart. Res. 2002, 4, 499-509.

35. Eliezer, S.; Eliaz, N.; Grossman, E.; Fisher, D.; Gouzman, I.; Henis, Z.; Pecker, S.; Horovitz, Y.; Fraenkel, M.; Maman, S.; Lereah, Y. Phy. Rev. 2004, 69, 144119/1-144119/6.

36. Ivanov, G. V.; Tepper, F. In Fourth International Symposium on Special Topics in Chemical Propulsion, Stockholm, 1997; p 636-644.

37. Phung, X.; Groza, J.; Stach, E. A.; Williams, L. N.; Ritchey, S. B. Mater. Sci. Eng. 2003, A359, 261-268.

38. Unal, A.; Leon, D. D.; Gurganus, T. B.; Hildemann, G. J. In ASM Handbook, 1998; Vol. 7.

39. Haber, J. A.; Buhro, W. E. J. Am. Chem. Soc. 1998, 120, 10847-10855.

40. Higa, K. T.; Johnson, C. E.; Hollins, R. A. US 3552946 1971, 1999.

41. Pesiri, D.; Aumann, C. E.; Bilger, L.; Booth, D.; Carpenter, R. D.; Dye, R.; O'Neill, E.; Shelton, D.; Walter, K. C. J. Pyrotech. 2004, 19, 19-31.

42. Gromov, A. A.; Foerter-Barth, U.; Teipel, U. Powder Technol. 2006, 164, 111-115.

43. Kwon, Y.-S.; Gromov, A. A.; Ilyin, A. P.; Rim, G.-H. App. Surf. Sci. 2003, 211, 57-67.

Page 60: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

50

44. Kwon, Y.-S.; Gromov, A. A.; Strokova, J. I. App. Surf. Sci. 2007, 253, 5558-5564.

45. Jouet, R. J.; Warren, A. D.; Rosenberg, D. M.; Bellitto, V. J.; Park, K.; Zachariah, M. R. Chem. Mater. 2005, 17, 2987-2996.

46. Fedotova, T. D.; Glotov, O. G.; Zarko, V. E. Propellants, Explos., Pyrotech. 2000, 25, 325-332.

47. Jouet, R. J.; Carney, J. R.; Granholm, R. H.; Sandusky, H. W.; Warren, A. D. Mater. Sci. Technol. 2006, 22, 422-429.

48. Zhukov, V.; Popova, I.; Fomenko, V.; Yates, J. T. Surf. Sci. 1999, 441, 240-250.

49. Zavyalov, S. A.; Pivkina, A. N.; Schoonman, J. Solid State Ionics 2002, 147, 415-419.

50. Zavyalov, S. A.; Timofeev, A. A.; Pivkina, A. N.; Schoonman, J. Nanostruct. Mater. 2002, 97-113.

51. Roy, C.; Dubois, C.; Lafleur, P.; Brousseau, P. Mater. Res. Soc. Sympos. Proc. 2003, 800, 79-84.

52. Brousseau, P.; Dubois, C. Mater. Res. Soc. Sympos. Proc. 2006, 896, 45-55.

53. Park, K.; Rai, A.; Zachariah, M. R. J. Nanopart. Res. 2006, 8, 455-464.

54. Park, K.; Lee, D.; Rai, A.; Mukherjee, D.; Zachariah, M. R. J. Phys. Chem. B 2005, 109, 7290-7299.

55. Foley, T. J.; Johnson, C. E.; Higa, K. T. Chem. Mater. 2005, 17, 4086-4091.

56. Shafirovich, E.; Bocanegra, P. E.; Chauveau, C.; Gokalp, I.; Goldshleger, U.; Rosenband, V.; Gany, A. Proc. Combust. Ins. 2005, 30, 2055-2062.

57. Shafirovich, E.; Mukasyan, A.; Thiers, L.; Varma, A.; Legrand, B.; Chauveau, C.; Gokalp, I. Combust. Sci. Technol. 2002, 174, 125-140.

Page 61: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

51

58. Yagodnikov, D. A.; Voronetskii, A. V. Fizika Goreniya i Vzryva 1997, 33, 60-68.

59. Kwon, Y.-S.; Gromov, A. A.; Ilyin, A. P. Combust. Flame 2002, 131, 349-352.

60. Shoshin, Y. L.; Mudryy, R. S.; Dreizin, E. L. Combust. Flame 2002, 128, 259-269.

61. Dreizin, E. L.; Shoshin, Y. L.; Mudryy, R. S. CPIA Publication 2001, 705, 189-199.

62. Pivkina, A.; Streletskii, A.; Kolbanev, I.; Ul'yanova, P.; Frolov, Y.; Butyagin, P.; Schoonman, J. J. Mater. Sci. 2004, 39, 5451-5453.

63. Streletskii, A. N.; Kolbanev, I. V.; Borunova, A. B.; Butyagin, P. Y. J. Mater. Sci. 2004, 39, 5175-5179.

64. Streletskii, A. N.; Kolbanev, I. V.; Borunova, A. B.; Leonov, A. V.; Butyagin, P. Y. Colloid J. 2004, 66, 729-735.

65. Streletskii, A. N.; Pivkina, A. N.; Kolbanev, I. V.; Borunova, A. B.; Leipunskii, I. O.; Pshechenkov, P. A.; Lomaeva, S. F.; Polunina, I. A.; Frolov, Y. V.; Butyagin, P. Y. Colloid J. 2004, 66, 736-744.

66. Krier, H.; Burton, R. L.; Pirman, S. R.; Spalding, M. J. J. Propul. Power 1996, 12, 672-679.

67. Pasternack, L. Combust. Flame 1992, 90, 259-268.

68. Pierson, H. O.; Mullendore, A. W. Thin Solid Films 1981, 83, 87-91.

69. Roberts, T. A.; Burton, R. L.; Krier, H. Combust. Flame 1993, 92, 125-143.

70. Macek, A.; Semple, J. M. Combust. Sci. Technol. 1969, 1, 181-91.

71. Foelsche, R. O.; Burton, R. L.; Krier, H. Combust. Flame 1999, 117, 32-58.

Page 62: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

52

72. Dreizin, E. L.; Keil, D. G.; Felder, W.; Vicenzi, E. P. Combust. Flame 1999, 119, 272-290.

73. Yetter, R. A.; Rabitz, H.; Dryer, F. L.; Brown, R. C.; Kolb, C. E. Combust. Flame 1991, 83, 43-62.

74. Brown, R. C.; Kolb, C. E.; Yetter, R. A.; Dryer, F. L.; Rabitz, H. Combust. Flame 1995, 101, 221-38.

75. Konczalla, M. Prog. Astronaut. Aeronaut. 1993, 152, 178-91.

76. Konczalla, M. Combust. 1993, 232-47.

77. Keil, D. G.; Dreizin, E. L.; Hoffman, V. K.; Calcote, H. F. Chem. Phys. Proc. Combust. 1999, 451-454.

78. Linder, D. P.; Page, M. J. Chem. Phys. 1995, 103, 8538-43.

79. Soto, M. R. J. Phys. Chem. 1995, 99, 6540-6547.

80. Politzer, P.; Lane, P.; Concha, M. C. Chem. Extreme Cond. 2005, 473-493.

81. Schoenitz, M.; Dreizin, E. L.; Shtessel, E. J. Prop. Power 2003, 19, 405-412.

82. Naslain, R. Prep. Methods Solid State Chem. 1972, 439-85.

83. Gutbier, A. Kolloid Z. 1914, 13, 137-143.

84. Elder, A. L.; Green, N. D. J. Phys. Chem. 1932, 36, 3085-6.

85. Weintraub, E. J. Indust. Eng. Chem. 1913, 5, 106.

86. Kiessling, R. Acta Chemica Scandinavica 1948, 2, 707-12.

Page 63: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

53

87. Laubengayer, A. W.; Hurd, D. T.; Newkirk, A. E.; Hoard, J. L. J. Am. Chem. Soc. 1943, 65, 1924-31.

88. Mellor, D. P.; Cohen, S. B.; Underwood, E. B. Aus. Chem. Ins. J. Proc. 1936, 3, 329-33.

89. Nisel'son, L. A.; Petrusevich, I. V.; Shamrai, F. I.; Fedorov, T. F. Zhurnal Prikladnoi Khimii 1962, 35, 984-9.

90. Casey, J. D.; Haggerty, J. S. J. Mater. Sci. 1987, 22, 737-744.

91. Steffens, K. L.; Zachariah, M. R.; DuFaux, D. P.; Axelbaum, R. L. Chem. Mater. 1996, 8, 1871-1880.

92. Johnston, H. L.; Hersh, H. N.; Kerr, E. C. J. Am. Chem. Soc. 1951, 73, 1112-1117.

93. Si, P. Z.; Zhang, M.; You, C. Y.; Geng, D. Y.; Du, J. H.; Zhao, X. G.; Ma, X. L.; Zhang, Z. D. J. Mater. Sci. 2003, 2003, 689-692.

94. Liu, H. C. Carbon 2000, 39, 144-147.

95. Kawaguchi, K. AIST Today (Japanese Edition) 2003, 3, 16.

96. Dubois, C.; Brousseau, P.; Roy, C.; Lafleur, P. AIChE Annual Meeting, Conference Proceedings, Austin, TX, United States, Nov. 7-12, 2004 2004, 281E/1-281E/9.

97. Pickering, A. L.; Mitterbauer, C.; Browning, N. D.; Kauzlarich, S. M.; Power, P. P. Chem. Commun. 2007, 580-582.

98. Slutsky, V. G.; Tsyganov, S. A.; Severin, E. S.; Polenov, L. A. Propellants, Explos., Pyrotech. 2005, 30, 303-309.

99. Dreizin, E. L.; Schoenitz, M.; Shoshin, Y. L.; Trunov, M. A.; Umbrajkar, S.; Ward, T. S.; Zhu, X. International Annual Conference of ICT 2005, 36th(Energetic Materials), 138/1-138/12.

100. Risha, G. A.; Boyer, E.; Evans, B.; Kuo, K. K.; Malek, R. Materials Research Society Symposium Proceedings 2003, 800, 243-254.

Page 64: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

54

101. Yang, Y.; Wang, S.; Sun, Z.; Dlott, D. D. Propellants, Explos., Pyrotech. 2005, 30, 171-177.

Page 65: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

55

Chapter 2: Synthesis, Characterization, and Functionalization of Boron Nanoparticles

Introduction

Elemental boron is a highly attractive fuel for propellants and explosives.1-12 Of all the

chemical elements, boron has the highest volumetric heat of combustion (33.4 kcal/cm3) and the

third highest gravimetric heat of combustion (14.1 kcal/g), after H2 and Be. These values are

over 3 times higher per unit volume, and 1.4 times higher per unit mass, than those of

hydrocarbon fuels.7,8

In order to increase the rate of energy release during the combustion of boron and make it

a more attractive fuel for propellants and explosives, it is advantageous to prepare it in a

nanoparticlulate form. In 1951 Johnson and coworkers13 prepared boron nanoparticles by the

thermolysis of diborane, B2H6, at 700 ºC using He as a carrier gas. The yield of the reaction is

quantitative and the particles are completely amorphous, with a size distribution of 25-500 nm

based on examination of electron micrographs. Several hundred grams of amorphous boron can

be been prepared by this method, but requires the use of explosive and toxic diborane.

In 1987 Casey et. al.14 synthesized boron nanoparticles via CO2 laser decomposition of

diborane or a mixture of boron trichloride and hydrogen gas. When the reduction of boron

trichloride was carried out at 930-1430 ºC, only 10-15% was reduced to boron nanoparticles.

The authors attributed the low yield of boron nanoparticles to the unusually high gas heating

rates (105-106 ºC sec-1), short resident times in the hot zone (0.001 seconds), and bypassing of

some of the gas mixture around the hot zone instead of through it. The boron nanoparticles

prepared in this method are amorphous as determined by powder X-ray diffraction experiments.

Four distinct types of particles are produced, as judged from the electron micrographs: small 20-

Page 66: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

56

30 nm particles, 80-100 nm perfect spheres, 150-200 nm non-spherical particles, and large

nanoparticles up to 1300 nm in size. The majority of the sample consists of the 150-200 nm non-

spherical particles, a finding that is also consistent with BET measurements. After being heated

to 1225 ºC in Ar, the boron nanoparticles were still amorphous as judged by powder X-ray

diffraction, but they did show electron diffraction peaks with d-spacings of 4.17, 2.37-2.46, 1.88-

1.97, 1.66-1.73, and 1.36-1.45 Å that correlate well with both β-rhombohedral and tetragonal

boron.

Casey and co-workers14 also carried out the laser assisted decomposition of diborane

using the same experimental setup. The optimal temperature for synthesizing boron

nanoparticles was 1460-1470 ºC, which produced boron nanoparticles in 83% yield with

diameters of 30-40 nm. The as synthesized boron nanoparticles were amorphous as determined

by powder X-ray diffraction, but crystallized upon heating to 1500 ºC in Ar. A 15-40 nm

crystallite size was determined from a Scherrer analysis. The d-spacings of 4.11 and 2.58 Å do

not appear to correlate with known boron phases. The crystallized boron nanoparticles oxidize

in the presence of oxygen at 400 ºC to give B2O3 as observed by X-ray powder diffraction.

Further heating of the sample in oxygen gave a B2O3 melt as observed by TGA/DTA. The

authors state the as synthesized boron nanoparticles are susceptible to aggregation as observed in

the electron micrographs.

More recently Si and coworkers15 reported the preparation of boron nanoparticles by arc

reduction of diborane at 2500 ºC. The particles prepared in this manner are 55-95 nm in size.

No powder X-ray diffraction experiments were conducted on the as synthesized boron

nanoparticles. Heating the particles under nitrogen at 1100 ºC affords BN particles.

Page 67: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

57

Pickering16 reported the synthesis of boron nanoparticles bearing bromine surface groups

by reduction of boron tribromide with sodium naphthalenide, and that the B-Br groups could be

converted to B-OR groups by treatment with alkoxides. Boron nanowires17,18 and thin films have

been grown by a variety of methods.

Decaborane, B10H14, is an air stable crystalline solid that is known to decompose on

surfaces at temperatures as low as 200 ºC to give boron films. Here we report the use of this

reagent as a starting material for the synthesis of boron nanoparticles. More significantly, we

report the first proof that the surfaces of boron nanoparticles can be derivatized, and made

amenable to further chemical manipulation.

Results and Discussion

Passage of decaborane vapor, with 1 atm argon as a carrier gas, through a hot zone at

700-900 ºC affords a grey-brown, non-pyrophoric powder. The powder consists of >97 % boron

and is free of hydrogen and carbon impurities as judged by combustion analysis. Typical

recovered yields are 40 %. The particles are easily suspended in organic solvents such as toluene

and pentane, and the resulting suspensions settle slowly over the course of a few hours. A

significant portion of the decaborane decomposes on the walls of the quartz tube yielding a black

film. Transmission electron microscopy images show that the majority of the particles have

diameters in the range 10 to 150 nm, and that they are textureless and approximately spherical

(Figure 2.1). Two broad peaks at d-spacings of 2.5 Å (2θ = 35.0º) and 4.0 Å (2θ = 22.5º) are

present in the Cu Kα XRD pattern (Figure 2.2). A Scherrer analysis of the powder X-ray

diffraction peak widths suggests that the crystalline domains are about 25 Å in size. The small

crystalline domain size is expected, because the nanoparticles are synthesized at temperatures far

Page 68: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

58

Figure 2.1. TEM image of boron nanoparticles supported on lacey carbon grids. The nanoparticles were prepared by thermolysis of decaborane at 700-900 °C under 1 atm Ar.

Page 69: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

59

Figure 2.2. Powder XRD spectrum of boron nanoparticles prepared by thermolysis of decaborane at 700-900 °C under 1 atm Ar.

Page 70: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

60

below the 2075 ºC melting point of boron.19 This result is in contrast to the particles formed by

the decomposition B2H6 or the reduction of boron trichloride by hydrogen gas at high

temperatures.15 The XRD pattern is very similar to that seen for commercial samples of boron

powder,20 and the d-spacings most closely resemble those of the α-rhombohedral phase.21,22 We

cannot rule out the possibility that amorphous boron may also be present.

The XPS spectrum (Figure 2.3) of the as-prepared particles show a signal at 189 eV (B

1s); this value is higher than the reported value owing to charging effects.23-26 Most notably,

there is no evidence of boron oxide. This result suggests that the as-synthesized boron particles

are relatively inert towards oxidation at room temperature. The (S)TEM-EELS spectrum of the

as-synthesized boron particles contains a signal at 210 eV that is also characteristic of pure boron

(Figure 2.4).

DSC experiments conducted on the boron nanoparticles under a pure oxygen atmosphere

show that an exothermic reaction begins at 440 ºC (Figure 2.5). The amount of heat released,

17.2 kJ/g, indicates that the oxidation of the nanoparticles is incomplete under these mild

conditions, and that only the surfaces of the particles are oxidized. (S)TEM-EELS studies of

boron nanoparticles heated in air for 4 hours at 600 ºC show that the oxidation produces a thin

oxide shell (~ 7 nm thick), with the core being pure boron (Figure 2.6).

In order to incorporate nanoparticles into larger assemblies and to stabilize them for long

term storage under ambient conditions, it is desirable to functionalize their surfaces. We find

that treatment of the boron nanoparticles with a ~6 mol percent benzene solution of XeF2 (a

convenient fluorinating agent) for 24-72 h, followed by removal of the solvent and excess XeF2

at 25 ºC under vacuum, yields boron particles whose surfaces are fluoride terminated. The

fluorinated particles are grey in color.

Page 71: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

61

 

Figure 2.3. XPS Spectrum of boron nanoparticles prepared by thermolysis of decaborane at 700-900 °C under 1 atm Ar. Oxygen and carbon impurities are due to the adhesive used to hold the nanoparticles in place.

Page 72: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

62

100 nm100 nm

Spectrum Image

Coun

ts

eV

0

2000

4000

6000

8000

10000

12000

14000

16000

18000

20000

Coun

ts

200 250 300 350 400 450 500 550 600 650 700eV

Figure 2.4. (S)TEM image of boron nanoparticles. EELS trace (along green line in image) showing contents of nanoparticles are pure boron. Signal at ~280 eV is due to carbon from the support. One data point was collected every 4.5 nm along the trace.

Page 73: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

63

Figure 2.5. DSC results for boron nanoparticles as-prepared (brown), surface-fluorinated (red), and surface-brominated (green).

Page 74: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

64

Figure 2.6. TEM image of a boron nanoparticle after oxidation in air at 600 ºC for 4 h. EELS trace (along red line in TEM image) showing incomplete oxidation of the boron nanoparticles.

Page 75: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

65

The XPS spectrum of the surfaces of the fluorinated particles shows signals at 189 eV (B

1s) and 691.0 eV (F 1s) due to bulk boron and to surface fluoride, respectively. As judged from

an XPS depth profile study (Figure 2.7), the fluorine atoms are confined to the surfaces of the

particles, whereas the interiors remain pure boron. The (S)TEM-EELS spectrum of the surfaces

of the boron nanoparticles treated with XeF2 show only the characteristic boron signal at 210 eV,

due to the low atomic concentration of fluorine present in the sample (<5%), but (S)TEM-EDS

experiments confirm the presence of fluorine with a characteristic fluorine signal at ~ 1 eV.

The TOF-SIMS spectra of the boron particles treated with XeF2 confirm the XPS

findings. In the negative ion scan (Figure 2.8), hydrogen, carbon, C-H, oxygen, O-H, and

fluorine are present. The hydrogen, carbon, C-H fragment, and O-H fragment can be attributed

to background gases in the instrument. In the positive ion scan (Figure 2.8), hydrogen, boron-10,

boron-11, carbon, and sodium are present in the spectrum. The hydrogen, carbon, and sodium

signals can all be attributed to background species. A depth profile analysis carried out by a

TOF-SIMS sputtering experiment (Ar ions) supports the conclusion that the fluorine is confined

to the surface of the particles, whereas the interiors of the particles are pure boron. The surfaces

of the boron particles can also be functionalized by treatment with neat bromine in benzene for

12 hours; removal of the benzene and excess bromine at 25 ºC under vacuum yields black boron

nanoparticles with a bromine content of 4.1%.

DSC experiments (Figure 2.5) conducted on the surface-fluorinated particles show that

they react exothermically under a pure oxygen atmosphere beginning at 460 ºC. The particles

treated with bromine (Figure 2.5) showed an even further increase in the onset temperature for

oxidation, with the exothermic reaction beginning at 540 ºC. The amount of heat released was

Page 76: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

66

 

Figure 2.7. XPS depth profile experiment (Ar ion sputtering) showing the decrease in the concentration of fluorine as a function of depth.

Page 77: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

67

0 5 10 15 20 250

500100015002000250030003500

Tota

l Cou

nts

(0.0

4 am

u Mass: 0.0342 Channel: 16517 Counts: 1 XEF2B_SCAN01_POSSEC.TDC + Ions 100µm 204545 cts

2312

10

1

11

0 10 20 30 400

100000

200000

300000

400000

500000

600000

Tota

l Cou

nts

(0.0

4 am

u bi

n)

Mass: 0.0426-0.0848 Channel: 16943-18555 Counts: 2 XEF2B_SCAN02_NEGSEC.TDC - Ions 100µm 1532357 cts

121713

1619

1

Figure 2.8. Top TOF-SIMS positive ion scan showing the presence of boron at masses 10 and 11. The signals at 1, 12, and 23 correspond to hydrogen, carbon, and sodium respectively. Bottom: TOF-SIMS negative ion scan showing the presence of fluorine at mass 19. The signals at 1, 12, 13, 16, and 17 correspond to hydrogen, carbon, a C-H fragment, oxygen, and a O-H fragment respectively.

Page 78: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

68

14.1 kJ/g and 16.9 kJ/g for the fluorinated and brominated boron nanoparticles, respectively,

again indicating that oxidation is incomplete.

Finally, we have reinvestigated the claim by Pickering16 that boron nanoparticles bearing

bromine surface groups can be prepared by reduction of boron tribromide with sodium

naphthalenide, and that the B-Br groups could be converted to B-OR groups by treatment with

alcohols. The resulting yellow oil has a 11B NMR chemical shift that is similar to that of

molecular B(OR)3 compounds. We find that the yellow oil consists predominantly of tri(n-

octyl)borate and n-octanol, not boron nanoparticles, as judged from microanalytical data (calc for

B(OC8H17)3: C, 72.4; H, 12.8. Found: C, 73.8; H, 14.2) and mass spectroscopy (calc for

B(OC8H17)3: 399. Found: 398.4). If elemental boron is present in the oil, it is a minor

constituent.

The present results show that highly pure boron nanoparticles can be synthesized by the

gas phase pyrolysis of decaborane with an argon carrier gas. The particle surfaces can be further

functionalized by fluorination or bromination. Thermal analysis studies of the particles show

that an exothermic combustion reaction takes place at temperatures above 500 ºC, and that

surface halogenation stabilizes the particles toward reaction with oxygen. (S)TEM-EELS

experiments show that the combustion of the particles is incomplete at these temperatures. .

Experimental Section

All experiments were carried out under vacuum or under argon by using standard

Schlenk or dry box techniques. The starting materials B10H14, Br2 (99.5+ %), and XeF2 were

used as received (Aldrich). Solvents were distilled from sodium-benzophenone (benzene,

pentane) or from sodium (toluene) and saturated with Ar before use. Microanalyses were

Page 79: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

69

performed by the Microanalytical Laboratory of the School of Chemical Sciences at the

University of Illinois.

TEM images were acquired either on a field emission JEOL 2010F (S)TEM instrument,

or on a JEOL 2010 LaB6 TEM. The XPS data were collected on a PHI 5400 instrument, and

XPS depth profiles were carried out by means of argon sputtering. TOF-SIMS measurements

were preformed on a PHI Trift III from Physical Electrons. X-ray diffraction data were collected

on a Bruker small angle X-ray scattering instrument in the George L. Clark X-ray facility at the

University of Illinois at Urbana-Champaign. TGA/DSC experiments were carried out at

Pennsylvania State by placing the samples into a ceramic crucible and collecting data on a

Q2010 TGA/DSC from TA Instruments. The heating rate was 20 K per minute and all

measurements were performed in an atmosphere of pure oxygen.

Synthesis of boron nanoparticles. A glass reservoir charged with decaborane (0.45 g,

4.0 mmol) was heated to 100 - 105 ºC and the vapor transported by means of an argon carrier (1

atm; 0.5 SCFH) through a vertical quartz tube maintained at 700 ºC or 900 ºC until all the

decaborane was consumed (6 h). The fine dark grey brown powder that accumulated downstream

of the hot zone was collected. Yield: 0.159 g (40 %). Anal: B, 97.2; C, 0.42; H, 0.0.

Fluorination of boron nanoparticles. Boron nanoparticles (0.100 g, 9.25 mmol)

suspended in benzene (20 mL) were treated with XeF2 (0.100 g, 0.591 mmol) dissolved in

benzene (20 mL). The mixture was stirred at room temperature for 72 h. Removal of the solvent

under reduced pressure yielded a grey solid. Yield: 40 mg (40%). Anal. B, 93.0; C, 0.81; H,

0.00; F, 1.30%.

Bromination of boron nanoparticles. Boron nanoparticles (0.115 g, 10.6 mmol)

suspended in benzene (20 mL) were treated with Br2 (3.12 g, 18.8 mmol). The mixture was

Page 80: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

70

stirred at room temperature for 12 h. Removal of the solvent and bromine under reduced

pressure at 30 º C yielded a black solid. Yield: 93.2 mg (85 %). Anal. B, 71.7; C, 4.04; H, 1.03;

Br, 4.12%.

Page 81: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

71

References

1. King, M. K. Combust. Sci. Technol. 1972, 5, 155-64.

2. Perchonok, E. In The Chemistry of Propellants; Penner, S. S., Ducarme, J., Eds.; Pergamon Press, 1960.

3. Natan, B.; Gany, A. J. Propul. Power 1993, 9, 694-701.

4. Karadimitris, A.; Scott, C.; Netzer, D.; Gany, A. J. Propul. Power 1991, 7, 341-7.

5. Safaneev, D. Z.; Kashporov, L. Y.; Grigorev, Y. M. Combustion, Explosion and Shockwaves 1981, 17, 210-214.

6. Natan, B.; Gany, A. J. Propul. Power 1981, 9, 155-157.

7. Natan, B.; Gany, A. J. Propul. Power 1991, 7, 37-43.

8. Gany, A.; Netzer, D. W. J. Propul. Power 1986, 2, 423-7.

9. Laredo, D.; Gany, A. Acta Astronaut. 1983, 10, 437-41.

10. King, M. K. J. Spacecr. Rockets 1982, 19, 294-306.

11. Merrill, K. K. Combust. Sci. Technol. 1974, 8, 255-273.

12. Gany, A.; Netzer, D. W. Int. J. Turbo and Jet Engines 1985, 2, 157-168.

13. Johnston, H. L.; Hersh, H. N.; Kerr, E. C. J. Am. Chem. Soc. 1951, 73, 1112-17.

14. Casey, J. D.; Haggerty, J. S. J. Mater. Sci. 1987, 22, 737-44.

Page 82: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

72

15. Si, P. Z.; Zhang, M.; You, C. Y.; Geng, D. Y.; Du, J. H.; Zhao, X. G.; Ma, X. L.; Zhang, Z. D. J. Mater. Sci. 2003, 38, 689-692.

16. Pickering, A. L.; Mitterbauer, C.; Browning, N. D.; Kauzlarich, S. M.; Power, P. P. Chem. Commun. 2007, 580-582.

17. Quandt, A.; Boustani, I. ChemPhysChem 2005, 6, 2001-2008.

18. Xu, T. T.; Nicholls, A. W.; Ruoff, R. S. NANO 2006, 1, 55-63.

19. Talley, C. P. J. Appl. Phys. 1959, 30, 1114-15.

20. Jiang, J.; Senkowicz, B. J.; Larbalestier, D. C.; Hellstrom, E. E. Supercond. Sci. Technol. 2006, 19, L33-L36.

21. Hoard, J. L.; Newkirk, A. E. J. Am. Chem. Soc. 1960, 82, 70-6.

22. Decker, B. F.; Kasper, J. S. Acta Cryst. 1959, 12, 503-6.

23. Ronning, C.; Schwen, D.; Eyhusen, S.; Vetter, U.; Hofsass, H. Surf. Coat. Technol. 2002, 158-159, 382-387.

24. Ennaceur, M. M.; Terreault, B. J. Nucl. Mater. 2000, 280, 33-38.

25. Ito, Y.; Arakawa, T.; Nishikawa, M. Plasma Sources Sci. Technol. 1996, 5, 305-310.

26. Moulder, J.; Stickle, W.; Sobel, P.; Bomben, E. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Division, Perkin-Elmer Corp.: Eden Prairie, 1995.

Page 83: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

73

Chapter 3: Nanoparticle Composites of Aluminum and Transition Metals. A Scanning

Transmission Electron Micrographic Study of Nanostructured Energetic Materials

Introduction

High energy nanomaterials such as nanothermites1-4 and Al nanoparticles5-8 have been a

topic of increasing research for applications as propellants and explosives.9-14 It is well known

that a strong correlation often exists between the energy density of a nanomaterial and its

sensitivity; a related issue is that nanoparticles are strongly driven to agglomerate and densify,

owing to their relatively high surface free energies.15 Considerable effort has been directed

towards the development of chemistries to decouple these phenomena and thus make the

materials safer and easier to handle.15 One approach to achieving this goal is to passivate the

nanomaterials by functionalizing their surfaces.16,17 The Al clusters of interest in this work

present an additional challenge related to their sensitivity to oxidative decomposition.18 An

important goal of this research is to identify and develop new methods that will generate Al

nanoclusters that are effectively stabilized against environmental degradation. These chemistries

must also preserve the value of the cluster as a high-energy additive material.

It has been shown that treating Al nanoparticles with long chain carboxylic acids19,20 or

perfluoroalkyl carboxylic acids21 increases the active aluminum content compared to

unpassivated Al nanoparticles, owing to the surface oxidation of the unpassivated particles. The

major drawback with utilizing organic capping groups is the lowering of weight percent active

Al compared to the content of organic material.

In 2005, Higa and co-workers reported the use of transition metals to inhibit the

formation of an oxide layer over the surface of aluminum nanoparticles.22 They claimed that

treatment of aluminum particles with solutions of certain transition metal complexes resulted in

Page 84: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

74

the deposition of a uniform coat of the zerovalent transition metal.22 These reactions, which take

advantage of the strongly reducing nature of metallic aluminum, may be termed redox

transmetalation reactions.23-27 The metal complexes are reduced to the zerovalent state, the

resulting transition metal atoms deposit on the surface, and at the same time the aluminum is

oxidized to Al3+, which dissolves in the solvent. The authors proposed that a thin layer of

amorphous transition metal coated the entire surface of the Al nanoparticle, although this

assertion has not been proven.

This study is intended to provide definitive information about the three-dimensional

composite architectures that these reactions actually generate. We used advanced analytical

electron microscopy techniques28 to image and differentiate the metallic components in atomistic

detail. We have found that instead of coating the entire surface of the aluminum nanoparticle

with a thin shell of transition metal, the aluminum nanoparticles are decorated with smaller

crystalline transition metal nanoparticles that form an incomplete shell.

Results and Discussion.

Synthesis and Bulk Characterization of Nanoparticle Composites of Aluminum and

Transition Metals. Aluminum particles were synthesized by thermolysis of

dimethylethylamine-alane in toluene/triethylamine, in the presence of a titanium(IV)

isopropoxide catalyst. Powder X-ray diffraction experiments showed that the as-prepared

(untreated) aluminum particles are crystalline, and give sharp peaks at 2θ = 38.5, 44.7, 65.1, and

78.2º that respectively correspond to the d-spacings associated with the (111), (002), (022) and

(113) planes in fcc aluminum.29 XPS spectra show the characteristic aluminum 2p signal at 73

eV.30

Page 85: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

75

The freshly prepared Al particles were then treated at room temperature under argon with

0.08 equiv of a transition metal complex in 1,2-dimethoxyethane, in order to deposit the

transition metals on the surfaces of the Al particles. The surface derivatizing agents studied were

the coinage metal complexes Cu(acac)2, Ag(acac), and AuCl(SMe2), and the nickel subgroup

complexes Ni(acac)2, Pd(acac)2, and Pt(acac)2, where acac = 2,4-pentanedionate. After a

reaction time of 17 h, the treated Al particles were collected, washed, and dried in vacuum. The

resulting particles, which were recovered in 70-90% yields, are highly reactive and

spontaneously combust in air.

Using the Al/Ni system as a case study, we determined how the final composition of the

nickel-treated Al particles depends on the reaction temperature, reaction time, and transition

metal complex loading. Somewhat surprisingly, these reaction parameters have very little

influence on the final compositions. Specifically, the nickel content is identical for Al particles

treated for 17 h with 0.08 equiv of Ni(acac)2 at 25 ºC, at 40 ºC, and at 80 ºC (refluxing dme).

Similarly, for treatment with 0.08 equiv of Ni(acac)2 at room temperature, the composition is

identical for reaction times of 1, 2.5, 3.5, 18, and 72 h. And finally, for reactions carried out for

17 hours at room temperature, the final compositions are essentially identical for loadings of

0.08, 0.5, 1.0, and 2.0 equivalents of the nickel complex.

Characterization of the Nanoparticle Composites. The transition metals are confined

to the surfaces of the particles and the particle interiors are still pure aluminum, as shown by

XPS experiments conducted with argon sputtering. Cs-STEM micrographs were acquired for

both as-prepared Al particles and particles that had been treated with the transition metal

complexes. Sample handling protocols were instituted to minimize atmospheric exposure, but

loading the samples into the instrument unavoidably entailed some exposure to air, and the

Page 86: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

76

results should be interpreted accordingly. Figures 3.1-3.6 show Al particles treated with copper,

silver, gold, nickel, palladium, and platinum complexes.

The as-prepared Al particles vary in size ranging from 30 to 500 nm, and their sizes were

essentially unchanged after treatment. In every case, however, treatment with the transition

metal complexes does not afford a uniform coating, but instead results in the formation of small

The sample grid was constructed of Mo.(Bottom Right) Size distribution histograms for platinum

supported on aluminum.

nanocrystals of zerovalent transition metals as disperse particles populating the surface of the

premade Al particles. These transition metal nanocrystals vary in size from atomic dispersions

to sizes exceeding 50 nm in diameter, depending on which transition metal was used. The

crystallinity of these transition metal deposits is shown by the clearly observable lattice fringes

and by their faceted shapes.

A magnified image of a Cu particle reveals an atomically resolved faceted crystal

structure complete with lattice fringes. For the Al/Cu sample (Figure 3.1), the EDS spectrum

collected at spot 1, a region between the appended smaller nanocrystals, shows no Cu signal but

contains strong signals for Al at 1.44-1.49 keV and 1.557 keV, energies corresponding to K-L1,2,3

and K-M1 excitations, respectively.31,32 The EDS spectrum of the Al surface also shows small

peaks for Ti (4.504-4.510 keV for the K-L2,3 transitions and 4.931 keV for K-M2,3 transitions),

which arises from the Ti-based catalyst used during sample preparation.32,33 Small

concentrations of Ti are present within all the Al particles. Conversely, spectra collected by

focusing the probe on one of the appended small nanocrystals (spot 2 in Figure 3.1) contain

peaks due to Cu at 1.022, 8.04, and 8.98 keV (for the L1-M2,3, K-L2,3, and K-M2,3,4,5 excitations,

respectively).32,33 Signals due to aluminum are also present in this spectrum; we believe that

Page 87: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

77

Figure 3.1. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with Cu(acac)2. (Top Right) Expanded view of boxed region in left image. (Bottom Left) EDS spectra of spots 1 and 2 in top left image. Spot 2 gives peaks due to Cu at 1.022, 8.04, and 8.98 keV due to the L1-M2,3, K-L2,3, and K-M2,3,4,5 excitations.32,33 The sample grid was constructed of Mo. (Bottom Right) Size distribution histogram of copper supported on aluminum.

Page 88: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

78

Figure 3.2. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with Ag(acac). (Top Right) Expanded view of boxed region in left image. (Bottom Left) EDS spectra of spots 1 and 2 in left image. Spot 2 gives Ag peaks at 2.633-3.525 keV due to L1,2,3 excitations to the M and N levels.32 The sample grid was constructed of Mo. (Bottom Right) Size distribution diagram of silver supported on aluminum.

Page 89: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

79

Figure 3.3. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with AuCl(SMe2). (Top Right) Expanded view of boxed region in left image. (Bottom Left) EDS spectra of spots 1 and 2 in left image. Spot 1 gives Au peaks at 9.713 and 10.308 keV for the L3-M5 and L2-M1 transitions, and a broad band enveloping the 11.371, 12.147, and 13.3-14.2 keV for the L1,2,3 transitions to the M and N levels.32 The sample grid was constructed of Mo. (Bottom Right) Size distribution histogram of gold particles supported on aluminum.

Page 90: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

80

Figure 3.4. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with Ni(acac)2. (Top Right) Expanded view of boxed region in left image. (Bottom Left) EDS spectra of spots 1 and 2 in left image. Spot 1 gives peaks due to Ni at 7.461-7.478 keV for K-L1,2,3 excitations and at 8.264-8.328 keV for K-M1,2,3,4,5 excitations.32,33 The sample grid was constructed of Mo. (Bottom Right) Size distribution histograms for nickel nanoparticles supported on aluminum.

Page 91: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

81

Figure 3.5. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with Pd(acac)2. (Top Right) Expanded view of boxed region in left image. (Bottom Left) EDS spectra of spots 1 and 2 in left image. Spot 2 gives peaks due to Pd between 2.833-3.533 keV for L1,2,3 excitations to the M and N levels.32 The sample grid was constructed of Mo. (Bottom Right) Size distribution histograms for palladium on aluminum.

Page 92: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

82

Figure 3.6. (Top Left) Cs-STEM micrograph of Al nanoparticles treated with Pt(acac)2. (Top Right) Expanded view of boxed region in left image arrows indicate small regions of increased contrast. (Bottom Left) EDS spectra of spots 1 and 2 in left image. Spot 2 gives peaks for Pt at 8.268 (L3-M1), 12.942 (L2-N5), and 13.156 keV (L1-M1), and broader bands spanning 9.362-9.442 (L3-M4,5), 11.044-11.250 (L1,2,3-M4,5 and N2,3,4), and 13.272-13.361 keV (L1-M2,3).32

Page 93: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

83

these signals do not indicate that a Cu/Al alloy or multicrystal has formed, but instead arise from

the Al support underlying the transition metal nanocrystal. The small sizes of these latter

crystals (4.3 ± 1.9 nm, Figure 3.1) make it difficult to obtain EDS signals from just the surface-

bound nanocrystals, a point further emphasized by the fact that 84 % of the Cu particles are

between 2 and 5.5 nm.

The powder diffraction data confirm the presence of Cu nanoparticles (Figure 3.7).

Mixing of the metal components was ruled out by the observed lattice constants of 3.615(2) Å

for Cu and 4.0444(3) Å for Al, which are very similar to their accepted values of 3.625 Å for

Cu34 and

4.04975 Å for Al.29 The Cu peaks are slightly broadened, and Scherrer analysis gives a particle

size of 9.5(5) nm.

Micrographs of the Al/Ag samples (Figure 3.2) reveal the presence of small Ag

nanocrystals that appear to be twinned. EDS confirmed that the low-contrast regions are Al

whereas the high-contrast regions are Ag (although the latter regions also show a signal from the

Al support).32 The Ag nanocrystals appear to be distributed randomly on the Al surface, with

some signs of aggregation being evident from inspection of multiple micrographs. These particle

sizes follow a Weibull size distribution (Figure 3.2), and the average Ag nanoparticle size, 5.7 ±

4.7 nm, is larger than seen for Cu on Al.

The diffraction data (Figure 3.8) are complicated by the fact that Al and Ag have the

same crystal structure and similar lattice parameters, so that the peaks due to these two species

overlap. The overlapping makes it difficult to apply the Scherrer equation to estimate the

average Ag particle size. Although the smaller angle peaks (at ~39 and ~45°) are dominated by

the reflections associated with Al, it is clear that on average the dimensions of the Ag particles

Page 94: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

84

Figure 3.7. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with copper nanoparticles, metallic aluminum (red line) calculated from the crystal structure, and

metallic copper35,36 (green line) calculated from the crystal structure.

Page 95: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

85

exceed a few nanometers.37-39 This conclusion is evidenced by the lack of broadening in the

partially resolved (022) and (113) Bragg peaks. The lattice parameters derived from a Rietveld

refinement are 4.083(3) and 4.0490(8) Å for Ag and Al, respectively,40 which compare well with

literature values for bulk Ag (4.0853 Å)41 and Al (4.04975 Å).29 We conclude that there is no

alloying between the Ag and Al.

Micrographs of the Al/Au samples show that the Au nanocrystals agglomerate on the Al

surface (Figure 3.3). Because of the larger disparity in nuclear charge, the contrast between the

Al surface and the Au nanoparticles is much higher than for the Al/Cu and Al/Ag samples. EDX

measurements identified and confirmed the presence of these Au clusters on the Al (Figure 3.3).

The dispersions range from sizes as small as 2.6 nm to as large as 68.5 nm. The size distribution

histogram (Figure 3.3) is relatively flat within this size regime, a fact reflected in the uncertainty

of the average particle size (15.6 ± 12.1 nm). As seen for the other two coinage metals, close

inspection of an atomically resolved Au particle reveals a series of stacked atomic planes.

The Bragg peaks for Au overlap with those of Al (Figure 3.9), as expected from the

similarity in their lattice constants of 4.072 Å35 and 4.04975 Å,29 respectively. Rietveld analysis

gave lattice constants of 4.074(4) and 4.056(7) Å that were incompatible with any known Au-Al

alloy, and suggest that the two metals have not mixed.42

The Al particles were also surface-derivatized with nickel, palladium, and platinum. For

the Al/Ni samples (Figure 3.4), the nickel nanoclusters on average were smaller in size (3.8 ± 1.7

nm; Figure 3.4) than any of the coinage metals. Like the Al/Cu specimen, the majority of the Ni

particles (81%) have sizes between 2 and 5 nm. The Ni nanocrystals show resolved atomic

Page 96: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

86

 

Figure 3.8. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with silver nanoparticles, metallic aluminum (red line) calculated from the crystal structure, and

metallic silver35,41,43 (green line) calculated from the crystal structure.

Page 97: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

87

structure, and there are indications that smaller Ni particles or atomic dispersions are present on

the Al surface (Figure 3.10). Diffraction peaks due to nickel were not observed in the powder X-

ray diffraction spectrum (Figure 3.11), probably because nickel was present in low atomic

concentrations (<5%) and has a low scattering cross section.44

In the Al/Pd composites, the Pd nanocrystals (Figure 3.5) vary significantly in size (8.8 ±

7.1 nm), and this behavior is similar to that seen for Al/Ag and Al/Au. A magnified region of the

boxed area in the micrograph shows that the Pd particles are agglomerates and contain multiple

crystal grains. The powder diffraction data confirms the presence of Pd nanoparticles as

determined by the broadened peak widths at angles satisfying the Bragg condition for Pd (Figure

3.12). Mixing of the metal components was ruled out by refinement of the data whereby it was

found that the lattice constant for the Pd data (3.8903 Å, and Al (4.0467 Å) for that matter) was

unaltered from its accepted value of 3.8902 Å.45

On the other hand, for the Al/Pt samples the Al particle is largely free of Pt nanocrystals

agglomerates (Figure 3.6). The high atomic scattering factor of Pt atoms allows them to be

distinguished above the low-Z Al background. A magnified image shows that ~1.5 nm molecular

clusters and individual Pt atoms (arrows) are sprinkled across the Al surface, in addition to the

~4 nm Pt nanocrystals. The molecular clusters constitute the majority of the deposited Pt content

as determined by particle size measurements. The mean particle size is 1.7 ± 1.6 nm with 73 %

of the particles having sizes between 0.5 – 2 nm. Of course, these measurements exclude the

contribution of free Pt atoms on the Al surface.

The X-ray diffractograms again contain low-intensity, broadened Bragg reflections

indicative of nanosized crystallites (Figure 3.13). The lattice parameter of 3.902 (6) Å is slightly

Page 98: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

88

different from that of bulk Pt (3.923 Å).45 This result can be explained by particles size effects,

which are known to affect lattice constants. 46

As has been documented in the past for Pt and Au particles,46 a contraction of the Pt lattice by

~0.53% is expected for the 1.5 nm particle sizes seen in the present study.

Deposition Mechanism. Previously, Higa and coworkers reported that aluminum

particles could be passivated by coating them with transition metals to form a kind of core-shell

structure. Our results show that Higa’s protocol – treating the Al particles with a solution of a

transition metal salt – does not result in the formation of a uniform coat, but instead generates 1-

5 nm nanocrystals of the zerovalent transition metal that are randomly distributed over the

surface of the aluminum particles.

The formation of decorated particles rather than core-shell structures is likely a

consequence of the deposition mechanism, which involves a redox transmetalation reaction

between zerovalent aluminum and a transition metal salt.47-49 Inevitably the surface of the

aluminum nanoparticles are passivated with a nonuniform coat of oxide. We propose that the

site for growth of the transition metal nanoparticles is in the defect zones where the aluminum

nanoparticle is not fully oxidized. The random growth of the transition metal nanocrystals could

be explained by the randomness of the oxide coat on the aluminum nanocrystal.

Page 99: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

89

 

Figure 3.9. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with gold nanoparticles, metallic aluminum (red line) calculated from the crystal structure, and

metallic gold43,50,51 (green line) calculated from the crystal structure.

Page 100: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

90

Figure 3.10. Magnified images of Ni nanoparticles on Al show small, high-intensity regions

above the Al background intensity.

Page 101: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

91

Figure 3.11. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with nickel nanoparticles, metallic aluminum (red line) calculated from the crystal structure, and

metallic nickel43,52,53 (green line) calculated from the crystal structure.

Page 102: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

92

 

Figure 3.12. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with palladium nanoparticles, metallic aluminum (red line) calculated from the crystal structure,

and metallic palladium45,54,55 (green line) calculated from the crystal structure.

Page 103: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

93

 

Figure 3.13. Powder XRD experimental data (black line) for aluminum nanoparticles decorated

with platinum nanoparticles, metallic aluminum (red line) calculated from the crystal structure,

and metallic platinum42,56 (green line) calculated from the crystal structure.

Page 104: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

94

Experimental Section

All reactions were performed using standard Schlenk techniques under argon unless

otherwise noted. Solvents were distilled either from sodium benzonephone ketal (pentane and

1,2-dimethoxyethane) or molten sodium (toluene, triethylamine). Dimethylethylamine alane (1.0

M solution in toluene), titanium isopropoxide, Ni(acac)2, Cu(acac)2, Au(dms)Cl, Ag(acac),

Pd(acac)2, and Pt(acac)2 were used as received (Aldrich). ICP analyses were performed by the

microanaltical laboratory at the University of Illinois at Urbana-Champaign. The powder

diffraction experiments were performed on a Bruker general area detector diffraction system

(GADDS) at the George L. Clark X-ray diffraction facility at the University of Illinois at

Urbana-Champaign. Samples were prepared for powder diffraction studies by packing 0.3 mm

capillaries in a glove box and sealing the tops with epoxy. After the samples were removed from

the glove box atmosphere, the capillaries were flamed sealed under Ar. XPS spectra were

collected on a PHI 5400 XPS instrument.

Aberration-corrected scanning transmission electron microscopy (Cs-STEM) images were

collected on a JEM 2200FS microscope operated at 200 keV with an electron probe capable of

resolving sub-Å features.2 The dark-field detector had an inner cut-off angle of 100 mrad to

minimize the collection of incoherently scattered electrons. The Z-contrast in the images was

increased by taking advantage of the dependence of the scattering angle of the coherently

scattered electrons on atomic number (~Z2), a technique often employed when imaging metallic

nanoscale catalysts supported on low-Z supports (e.g. C, Al2O3, etc.).57 The microscope was

fitted with an INCA Energy 200Premium EDX-System equipped with a Si(Li) super-

atmospheric thin window detector (Oxford Instruments) for energy dispersive X-ray

Page 105: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

95

spectroscopy (EDS) measurements. Samples of the Al/Cu particles were loaded onto Mo grids

(Mo-200LC, Pacific Grid Tech) and the grids were mounted onto a beryllium double-tilt holder.

■ Caution: Aluminum nanoparticles are highly pyrophoric, and should be handled under

an inert atmosphere at all times. An effective procedure for disposing of the nanoparticles is to

place a septum in one of the ground glass joints of the container, and to allow atmospheric

oxygen to diffuse slowly into the flask through a needle placed in the septum. After five hours

the samples are usually sufficiently oxidized that they can be slowly hydrolyzed under Ar with

water, followed by 2.0 M hydrochloric acid. After this treatment the samples can be disposed in

solid waste streams.

Aluminum Nanoparticles. These were prepared by a modification of a literature

recipe.22 To titanium isopropoxide (1.48 g, 5.2 mmol) in toluene (20 mL) was added

triethylamine (10.0 mL, 71 mmol). The solution was heated to 80 ºC and stirred for five min.

Dimethylethylamine alane (100 mL of a 1 M solution in toluene, 100 mmol) was added dropwise

to the hot solution over 20 min. Gas was evolved and the solution changed color immediately

from clear to black. A gray solid precipitated and aluminum mirror formed on the sides of the

flask. The reaction mixture was stirred for 4 h at 80 ºC and then was cooled in an ice bath. The

black powder was collected by filtration, washed with toluene (2 × 25 mL) and pentane (3 × 35

mL), and dried in vacuum. Yield: 2.15 g (80 %). Anal. Found: Al, 95.1 %, Ti, 3.41 %.

General procedure for reactions of aluminum nanoparticles with metal salts. To a

suspension of freshly prepared aluminum nanoparticles (0.46 g, 17 mmol) in 1,2-

dimethoxyethane (10 mL) was added dropwise over 20 min a solution Ni(acac)2 (0.34 g, 1.35

mmol) in 1,2-dimethoxyethane (25 mL). The resulting dark slurry was stirred for 17 h, after

which the black powder was collected by filtration, washed with 1,2-dimethoxyethane (3 × 15

Page 106: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

96

mL) and pentane (3 × 25 mL), and dried in vacuum. Typical recovery yields are 70-90% of the

initial mass of aluminum.

Page 107: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

97

References

1. Wilson, D. E.; Schroder, K. A.; Willauer, D. L.; Bless, S.; Russell, R. T. Patent Application: US 2009301337.

2. Apperson, S. J.; Bezmelnitsyn, A. V.; Thiruvengadathan, R.; Gangopadhyay, K.; Gangopadhyay, S.; Balas, W. A.; Anderson, P. E.; Nicolich, S. M. J. Propul. Power 2009, 25, 1086-1091.

3. Puszynski, J. A. J. Therm. Anal. Calorim. 2009, 96, 677-685.

4. Johnson, C. E.; Higa, K. T.; Albro, W. R. Proc. Int. Pyrotech. Semin. 2008, 35th, 159-168.

5. Muravyev, N.; Frolov, Y.; Pivkina, A.; Monogarov, K.; Ivanov, D.; Meerov, D.; Fomenkov, I. Cent. Eur. J. Energ. Mater. 2009, 6, 195-210.

6. Hahma, A. Patent Application: WO 2009127784.

7. Jayaraman, K.; Anand, K. V.; Bhatt, D. S.; Chakravarthy, S. R.; Sarathi, R. J. Propul. Power 2009, 25, 471-481.

8. Umbrajkar, S. M.; Seshadri, S.; Schoenitz, M.; Hoffmann, V. K.; Dreizin, E. L. J. Propul. Power 2008, 24, 192-198.

9. Pantoya, M. L.; Hunt, E. M. Appl. Phys. Lett. 2009, 95, 253101/1-253101/3.

10. Gromov, A.; Strokova, Y.; Kabardin, A.; Vorozhtsov, A.; Teipel, U. Propellants, Explos., Pyrotech. 2009, 34, 506-512.

11. Sinditskii, V. P.; Egorshev, V. Y.; Serushkin, V. V.; Levshenkov, A. I.; Berezin, M. V.; Filatov, S. A.; Smirnov, S. P. Thermochim. Acta 2009, 496, 1-12.

12. Bhattacharya, A.; Guo, Y. Q.; Bernstein, E. R. J. Chem. Phys. 2009, 131, 194304/1-194304/8.

Page 108: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

98

13. Li, H.; Meziani, M. J.; Lu, F.; Bunker, C. E.; Guliants, E. A.; Sun, Y.-P. J. Phys. Chem. C 2009, 113, 20539-20542.

14. Cheng, J.; Hng, H. H.; Ng, H. Y.; Soon, P. C.; Lee, Y. W. J. Mater. Res. 2009, 24, 3220-3225.

15. Fenski, D.; Longoni, G.; Schmid, G. Clusters and Colloids: From Theory to Applications; VCH: New York, 1994.

16. Parker, A. J.; Childs, P. A.; Palmer, R. E.; Brust, M. App. Phys. Lett. 1999, 74, 2833-2835.

17. Brust, M.; Fink, J.; Bethell, D.; Schiffrin, D. J.; Kiely, C. Chem. Commun. 1995, 1655-1656.

18. Sanchez-Lopez, J. C.; Gonzalez-Elipe, A. R.; Fernandez, A. J. Mater. Res. 1998, 13, 703-710.

19. Lyashko, A. P.; Il'in, A. P.; Savel'ev, G. G. Zh. Prikl. Khim. 1993, 66, 1230-3.

20. Kwon, Y.-S.; Gromov, A. A.; Ilyin, A. P.; Rim, G.-H. Appl. Surf. Sci. 2003, 211, 57-67.

21. Jouet, R. J.; Warren, A. D.; Rosenberg, D. M.; Bellitto, V. J.; Park, K.; Zachariah, M. R. Chem. Mater. 2005, 17, 2987-2996.

22. Foley, T. J.; Johnson, C. E.; Higa, K. T. Chem. Mater. 2005, 17, 4086-4091.

23. Pasricha, R.; Bala, T.; Biradar, A. V.; Umbarkar, S.; Sastry, M. Small 2009, 5, 1467-1473.

24. Bala, T.; Enoki, T.; Prasad, B. L. V. J. Nanosci. Nanotechnol. 2007, 7, 3134-3139.

25. Ankamwar, B.; Damle, C.; Ahmad, A.; Sastry, M. J. Nanosci. Nanotechnol. 2005, 5, 1665-1671.

26. Selvakannan, P. R.; Sastry, M. Chem. Commun. 2005, 1684-1686.

Page 109: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

99

27. Bala, T.; Arumugam, S. K.; Pasricha, R.; Prasad, B. L. V.; Sastry, M. J. Mater. Chem. 2004, 14, 1057-1061.

28. Sanchez, S. I.; Small, M. W.; Sivaramakrishnan, S.; Wen, J.-g.; Zuo, J.-M.; Nuzzo, R. G. Anal. Chem. 2010, 82, 2599-2607.

29. Cooper, A. S. Acta. Crystallogr. 1962, 15, 578-583.

30. Moulder, J.; Stickle, W.; Sobel, P.; Bomben, E. Handbook of X-Ray Photoelectron Spectroscopy; Physical Electronics Division, Perkin-Elmer Corp.: Eden Prairie, 1995.

31. Schweppe, J.; Deslattes, R. D.; Mooney, T.; Powell, C. J. J. Electron Spectrosc. Relat. Phenom. 1994, 67, 463-78.

32. Bearden, J. A. Rev. Mod. Phys. 1967, 39, 78-124.

33. Holzer, G.; Fritsch, M.; Deutsch, M.; Hartwig, J.; Forster, E. Phys. Rev. A: At., Mol., Opt. Phys. 1997, 56, 4554-4568.

34. Straumanis, M. E.; Yu, L. S. Acta Crystallogr., Sect. A 1969, 25, 676-82.

35. Suh, I. K.; Ohta, H.; Waseda, Y. J. Mater. Sci. 1988, 23, 757-60.

36. Straumanis, M. E.; Yu, L. S. Acta Crystallogr., Sect. 1969, 25, 676-82.

37. Debye, P. Ann. Phys. 1915, 46, 809-23.

38. Scherrer, P. Göttinger Nachrichten Gesell. 1918, 2, 98.

39. Patterson, A. L. Phys. Rev. 1939, 56, 978-82.

40. Rietveld, H. M. J. Appl. Crystallogr. 1969, 2, 65-71.

41. Liu, L. G.; Bassett, W. A. J. Appl. Phys. 1973, 44, 1475-9.

Page 110: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

100

42. Owen, E. A.; Yates, E. L. Philos. Mag. 1933, 15, 472-88.

43. Jette, E. R.; Foote, F. J. Chem. Phys. 1935, 3, 605-16.

44. McMaster, W. H.; Del Grande, N. K.; Mallett, J. J.; Hubbell, J. H. Compilation of X-Ray Cross Sections, Lawerence Livermore Lab., Report UCRL-50174 1969.

45. Schroeder, R. H.; Schmitz-Pranghe, N.; Kohlhaas, R. Z. Metallk. 1972, 63, 12-16.

46. Solliard, C.; Flueli, M. Surf. Sci. 1985, 156, 487-94.

47. Lin, W.; Nuzzo, R. G.; Girolami, G. S. J. Am. Chem. Soc. 1996, 118, 5988-5996.

48. Lin, W.; Wiegand, B. C.; Nuzzo, R. G.; Girolami, G. S. J. Am. Chem. Soc. 1996, 118, 5977-5987.

49. Crane, E. L.; You, Y.; Nuzzo, R. G.; Girolami, G. S. J. Am. Chem. Soc. 2000, 122, 3422-3435.

50. Salamakha, L. P.; Bauer, E.; Mudryi, S. I.; Goncalves, A. P.; Almeida, M.; Noel, H. J. Alloys Compd. 2009, 479, 184-188.

51. Maeland, A.; Flanagan, T. B. Can. J. Phys. 1964, 42, 2364-6.

52. Rouquette, J.; Haines, J.; Fraysse, G.; Al-Zein, A.; Bornand, V.; Pintard, M.; Papet, P.; Hull, S.; Gorelli, F. A. Inorg. Chem. 2008, 47, 9898-9904.

53. Lu, S. S.; Chang, Y. L. Proc. Phys. Soc., London 1941, 53, 517-28.

54. King, H. W.; Manchester, F. D. J. Phys. F 1978, 8, 15-26.

55. Hull, A. W. Phys. Rev. 1921, 17, 571-88.

56. Edwards, J. W.; Speiser, R.; Johnston, H. L. J. Appl. Phys. 1951, 22, 424-8.

Page 111: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

101

57. Varela, M.; Lupini, A. R.; van Benthem, K.; Borisevich, A. Y.; Chisholm, M. F.; Shibata, N.; Abe, E.; Pennycook, S. J. Annu. Rev. Mater. Res. 2005, 35, 539-569.

Page 112: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

102

Chapter 4. Synthesis of New Aminodiboranate Salts and Their 1,4-Dioxane Adducts

Introduction

Chemical vapor deposition (CVD) can be used to deposit high-quality thin films of metal

borides from single source molecular precursors. Some examples of the precursors used to

deposit metal boride films are the tetrahydroborate and octahydrotriborate complexes

Ti(BH4)3(dme),1-3 Zr(BH4)4,4 Hf(BH4)4,5-10 Cr(B3H8)2,11,12 and Mg(B3H8)2(Et2O)2.13 In many

cases the precursor used to deposit metal boride films can also be used to deposit thin films of

metal oxides, by performing the deposition in the presence of water.14

Only five homoleptic tetrahydroborate complexes are known to be monomers in the solid

state, those of Al,15,16 Zr,17 Hf,18,19 Np,20,21 and Pu.21 Homoleptic tetrahydroborate complexes of

the alkali metals,22-25 the alkaline earth metals,26 Ti,27,28 Th,28 Pa,21 and U17 are also known, but

exist as oligomers or polymers in the solid state. The rarity of homoleptic monomeric

tetrahydroborate complexes can be attributed to the small steric size of this borohydride ligand.

Using higher borohydrides has enabled the synthesis of a very small number of additional

homoleptic hydroborate complexes such as Mg(B3H8)2 and Cr(B3H8)2, and also a somewhat

larger number of heteroleptic complexes, although the latter are rarely volatile.11,13,29-40

In 1969 Keller reported the synthesis of the first aminodiboranate salt Na[(H3B)2NMe2]

by treatment of B2H5NMe2 with sodium hydride in 1,2-dimethoxyethane.41 The adduct

Na[(H3B)2NMe2](diox) was prepared by crystallization of the salt in the presence of 1,4-

dioxane.41 Nöth reported a new synthesis of Na[(H3B)2NMe2] in 1999: the sodium reduction of

dimethylamine borane in refluxing thf.42 Nöth also reported the preparations of

Na[(H3B)2NMe2](thf) and Na[(H3B)2NMe2](benzo-15-crown-5) by crystallization of the salt in

the presence of the corresponding ether.42

Page 113: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

103

The aminodiboranate ligand consists of two borohydride groups joined by an amido

linker, so that it is sterically larger than BH4- and B3H8

- and thus able to occupy a larger fraction

of a metal’s coordination sphere. Our group has recently prepared complexes of a variety of

alkaline earth, transition metal, lanthanide, and actinide complexes with the aminodiboranate

ligand (H3B)2NMe2-. We have also reported the synthesis of several new aminodiboranate

ligands in which the substituents on the nitrogen have been varied, which enables tuning of the

volatility and reactivity of the resulting metal complexes.43,44 Here we report the synthesis of

two new aminodiboranate salts in order to further our understanding of this truly remarkable

family of ligands.

Results and Discussion

Synthesis of New Sodium Aminodiboranate Salts and their 1,4-Dioxane Adducts.

The sodium reduction of diethylamine-borane and piperidine-borane in refluxing tetrahydrofuran

affords the new aminodiboranate salts Na[(H3B)2NEt2], 1, and Na[(H3B)2NC5H10], 2. The 11B

NMR spectra of these compounds consist of a binomial quartet at -14.8 (1JBH = 90 Hz) and δ -

13.5 (1JBH = 90 Hz), respectively. The 11B NMR spectra of the crude reaction solutions reveal

that a sodium borohydride (a pentet at δ -43.2 with 1JBH = 81 Hz, ~10 mol percent compared to 1

and 2) byproduct is formed. In the reaction that generates 1, a third byproduct is formed:

sodium octahydrotriborate, which appears as a septet at δ -31.0 with 1JBH = 33.5 Hz.45 The

sodium borohydride impurity can be removed by extracting the reaction products with diethyl

ether, in which only the aminodiboranate salts are soluble. The purified sodium aminodiboranate

salts can be handled in air for a few minutes, but must be stored under dry conditions.

Page 114: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

104

Adding 1,4-dioxane to 1 and 2 affords the adducts Na[(H3B)2NEt2](diox), 3, and

Na[(H3B)2NC5H10](diox)·diox, 4, respectively.

NMR Spectra of the Sodium Aminodiboranate Salts and their 1,4-Dioxane Adducts.

The 1H NMR spectrum of Na[(H3B)2NEt2], 1, in d8-thf shows three signals: a 1:1:1:1 quartet for

the borohydride hydrogens at δ 1.05 (1JBH = 90 Hz), a triplet for the methyl hydrogens at δ 1.03

(3JHH = 7.3 Hz), and a quartet for the methylene hydrogens at δ 2.45 (3JHH = 7.3 Hz). The 11B

NMR spectrum of Na[(H3B)2NEt2] in d8-thf shows a binomial quartet at δ -14.8 (1JBH = 90 Hz),

and the 13C{1H} NMR spectrum in d8-thf shows two resonances δ 53.6 and 10.4 for the α and β

carbons of the ethyl groups, respectively. The NMR spectra of the dioxane adduct 3 in d8-thf are

very similar to those of 1 (see Experimental Section), although small differences suggest that the

dioxane molecules are not being displaced by the d8-thf solvent.

The 1H NMR spectrum of Na[(H3B)2NC5H10], 2, in d8-thf shows a 1:1:1:1 quartet at δ

1.34 (1JBH = 90 Hz) for the borohydride hydrogens, and three multiplets for the piperidinyl

hydrogens. The 11B NMR spectrum of 2 in d8-thf shows a binomial quartet at δ -13.5 (1JBH = 90

Hz) and the 13C{1H} NMR spectrum shows three resonances for the piperidinyl carbons at δ

22.6, 25.5, and 58.9. Again, the NMR chemical shifts of the analogous dioxane adduct 4 in d8-

thf are nearly identical to these values.

Crystal Structures of the 1,4-Dioxane Adducts of the Sodium Aminodiboranate

Salts. The crystal structure of Na[(H3B)2NEt2](diox), 3, is shown in figure 4.1. The compound

crystallizes in the space group C2/c with one formula unit per asymmetric unit. Each sodium

atom is coordinated to two oxygen atoms of two different dioxane ligands and to one chelating

Page 115: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

105

Figure 4.1. ORETP representation of Na[(H3B)2NEt2](diox), 3. The 35% probability density

surfaces are shown, except for hydrogen atoms which are represented by arbitrarily sizes

spheres. The structure has been trimmed from an infinite chain to the smallest repeating unit and

all hydrogen atoms attached to carbon have been removed for clarity.

Page 116: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

106

aminodiboranate ligand, which is coordinated in a κ2H,κ2H fashion so that four hydrogen atoms

form contacts with the sodium cation. The Na[(H3B)2NEt2] units are arranged in pairs centered

on inversion centers; within each pair, three hydrogen atoms from one BH3 group bridge to the

other sodium atom. Each sodium atom therefore has an overall coordination number of 9: seven

hydrogen atoms and two oxygen atoms form the inner coordination sphere. One of the hydrogen

atoms on each (H3B)2NEt2 group is bound to both sodium atoms. If each BH3 group is regarded

as occupying one coordination site, then the coordination geometry about each Na center can be

viewed as a distorted trigonal bipyramid in which with O(1), O(2), and B(1) comprise the three

equatorial groups, and B(1A) and B(2) are the axial ligands. Each Na2[(H3B)2NEt2]2 pair is

linked each to four other such pairs by means of 1,4 dioxane molecules that bridge between

sodium atoms. This bridging interactions link the pairs into a three-dimensional polymeric

network.

The Na···B distances within the chelating interaction are 2.805(3) and 2.926(4) Å; these

distances are characteristic of κ2H interactions. In contrast, the 2.682(2)Å Na···B distance that

links the Na[(H3B)2NEt2] units within each pair is characteristic of a κ3H interaction. The

shortest Na-H contact of 2.37(2) Å occurs for a hydrogen atom that bridges between B and only

one Na center; several of the hydrogen atoms bridge between B and two Na centers, and these

contacts are longer. The Na(1)-O(1) and Na(1)-O(2) distances are 2.380(1) and 2.367(1) Å,

respectively.

The compound Na[(H3B)2NC5H10](diox)·diox, 4, also crystallizes in the space group

C2/c with one formula unit (which includes an unligated dioxane molecule) in the asymmetric

unit. As in 3, each sodium atom is coordinated to two oxygens of two different dioxane ligands

and to one κ2H,κ2H chelating aminodiboranate ligand (Figure 4.2). The Na[(H3B)2NC5H10] units

Page 117: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

107

Figure 4.2. ORETP representation of Na[(H3B)2NC5H10](diox)·diox, 4. The 35% probability

density surfaces are shown, except for hydrogen atoms which are represented by arbitrarily sizes

spheres. The structure has been trimmed from an infinite chain to the smallest repeating unit.

Page 118: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

108

are arranged in pairs, in which one BH3 group forms a κ3H interaction with the second sodium

center, as seen in 3 and in the previously reported structure of Na[(H3B)2NC4H8](diox).44 If the

BH3 groups are regarded as occupying one coordination site, the geometry about each Na atom is

again a distorted trigonal bipyramid with O(1), O(2) and B(1) in the equatorial sites and B(2) and

B(1A) in the axial sites. The dioxane molecules link the Na2[(H3B)2NC5H10]2 pairs into a

polymeric network. As before, the shortest Na···B distance of 2.707(4) Å is associated with the

κ3H interactions that link the two Na[(H3B)2NC5H10] units into pairs, and the two Na···B

distances involving the κ2H chelating interaction of and 2.872 (4) Å are somewhat longer. The

Na(1)-O(1) and Na(1)-O(2) distances of 2.341(2) Å and 2.307 (2) Å are slightly different. The

shortest Na-H contacts of 2.41(3) Å is seen for a hydrogen atom that bridges between B and one

Na atom; longer contacts are seen for hydrogen atoms that interact with two Na atoms.

Experimental Section

All operations were carried out under argon or vacuum unless otherwise specified.

Solvents were distilled under nitrogen either from sodium and benzophenone (thf, ether, pentane,

and benzene) or from molten sodium (1,4-dioxane). Diethylamine, piperidine, 1.0 M solution of

BH3 in thf, and sodium metal were purchased from Aldrich and used as received.

The IR spectra were recorded on a Nicolet Impact 410 FT-IR instrument as Nujol mulls

between KBr plates. Elemental analyses were performed by the University of Illinois

Microanalytical Laboratory. 1H NMR spectra were recorded on a Varian Unity Inova-500NB at

11.75 T, 13C{1H} NMR spectra were recorded on a Varian Unity-500 spectrometer at 11.75 T,

and 11B and 11B{1H} NMR spectra were recorded on a Varian Unity Inova-400 at 9.4 T. NMR

Page 119: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

109

chemical shifts are reported in δ units (positive chemical shifts to higher frequencies relative to

SiMe4 or BF3·Et2O).

General Preparation of Borane Amine Adducts. To BH3·thf (200 mL of a 1 M

solution in tetrahydrofuran, 0.200 mol) at 0 ºC was added the dry amine (0.200 mol) dropwise

over 20 min. The resulting solution of the amine-BH3 reagent was used directly in the following

reactions. The 11B NMR spectra of thf solutions of C5H10N·BH3 and Et2HN·BH3 consist of a

binomial quartet at -14.6 (1JBH = 96 Hz) and -16.6 (1JBH = 97 Hz), respectively.

Sodium N,N-Diethylaminodiboranate, Na[(H3B)2NEt2], 1. A 500 mL round bottomed

flask was equipped with a large stirring bar and approximately fifteen 1 × 1 × 1 cm cubes of

freshly cut sodium (14 g, 0.61 mol). A solution of diethylamine-BH3 (0.200 mol) in

tetrahydrofuran (200 mL) was added dropwise over 20 min. No gas evolution occurred. After

the addition was complete, the round bottomed flask was fitted with a water cooled condenser

and the solution was heated to reflux until all of the starting material had been consumed (as

monitored by 11B NMR spectroscopy). Typical reaction times are 70-80 h. The reaction mixture

was cooled and filtered. The colorless filtrate was taken to dryness under vacuum to give a sticky

grey-white residue. The residue was washed with benzene (2 × 120 mL) and pentane (3 × 50

mL) and dried under vacuum to yield a free flowing white powder. Yield: 4.68 g (38%). Anal.

Calc for C4H16NB2Na: C, 39.1; H, 13.1; N, 11.4. Found: C, 39.0; H, 13.6; N, 11.1. 1H NMR (d8-

thf): δ 1.03 (t, 3JHH = 7.3 Hz, β-Et, 6H), 1.05 (1:1:1:1 q, 1JBH = 90 Hz, BH3, 6H), 2.45 (q, 3JHH =

7.3 Hz, α-Et, 4H). 13C{1H} NMR (d8-thf): δ 10.4 (s, β-Et), 53.6 (s, α-Et). 11B NMR (d8-thf): δ -

14.8 (q, 1JBH = 90 Hz). IR (cm-1): 2326 s br, 2256 s br, 1217 m br, 1164 m br, 1151 m br, 1130

m, 1049 m, 972 w, 845 w, 800 w, and 779 w.

Page 120: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

110

Sodium Piperidinyldiboranate, Na[(H3B)2NC5H10], 2. A 500 mL round bottomed

flask was equipped with a large stirring bar and approximately fifteen 1 × 1 × 1 cm cubes of

freshly cut sodium (14 g, 0.61 mol). A solution of pipiderine-BH3 (0.200 mol) in

tetrahydrofuran (200 mL) was added dropwise over 20 min. Gas evolution occurred

(presumably H2), but the solution remained colorless. After the addition was complete, the

round bottomed flask was fitted with a water cooled condenser, and the solution was heated to

reflux until all of the starting material had been consumed (as monitored by 11B NMR

spectroscopy). Typical reaction times were 40-50 h. The reaction mixture was cooled and

filtered to give a clear colorless solution. The filtrate was taken to dryness under vacuum to

yield a sticky grey-white residue, which was washed with benzene (2 × 120 mL) and pentane (3

× 50 mL) and then dried under vacuum to afford a white solid. Yield: 6.25 g (46%). Anal. Calc

for C5H16NB2Na: C, 44.6; H, 12.0; N, 10.4. Found: C, 44.5; H, 12.4; N 10.3. 1H NMR (d8-thf):

δ 1.34 (1:1:1:1 q, 1JBH = 90 Hz, BH3, 6H), 1.36 (pentet, 3JHH = 5.7 Hz, 4-CH2, 2H), 1.61 (m, 3-

CH2, 4H), 2.48 (m, 2-CH2 , 4H). 13C{1H} MR (d8-thf): δ 22.6 (s, 4-CH2), 25.5 (s, 3-CH2), 58.9

(s, 2-CH2). 11B NMR (d8-thf): δ -13.5 (q, 1JBH = 90 Hz). IR (cm-1): 2326 s br, 2256 s br, 1217

m br, 1164 m br, 1130 m, 1049 m, 972 m, 846 w, and 669 w.

Sodium N,N-Diethylaminodiboranate 1:1 Dioxane, Na[(H3B)2NEt2](diox), 3. To

Na[(H3B)2NEt2] (0.100 g, 0.814 mmol) was added 1,4-dioxane (5.00 mL, 58.7 mmol). After the

slurry had been stirred for 20 min, the solvent was removed under vacuum and the white residue

was extracted with diethyl ether (2 × 20 mL). The extract was filtered, and the colorless filtrate

was concentrated to ca. 5 mL and cooled to -20 ºC to afford large colorless needles. Yield:

0.0638 g (37 %). Anal. Calc for C8H24NO2B2Na: C, 45.6; H, 11.5; N, 6.64. Found: C, 45.6; H,

11.5; N, 6.69. 1H NMR (d8-thf): δ 1.03 (t, 3JBH = 7.3 Hz, β-Et, 6H), 1.18 (1:1:1:1 q, 1JBH = 92

Page 121: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

111

Hz, BH3, 6H), 2.48 (q, 3JHH = 7.3 Hz, α-Et , 4H), 3.56 (s, diox, 8H). 13C{1H} NMR (d8-thf): 10.5

(s, β-Et), 53.7 (s, α-Et), 67.8 (s, diox). 11B NMR (d8-thf): δ -14.6 (q, 1JBH = 92 Hz). IR (cm-1):

2397 m br, 2295 br, 2249 br, 1637 m, 1617 m, 1377 s, 1295 m, 1258 m, 1195 br, 1164 br, 1149

br, 1116 br, 1084 w, 1050 m, 960 m, 889 m, 875 m, 798 w, 773 m, and 615 w.

Sodium Piperidinyldiboranate 1:2 Dioxane, Na[(H3B)2NC5H10](diox)·diox, 4. To

Na[(H3B)2NC5H10] (0.100 g, 0.742 mmol) was added 1,4-dioxane (5.00 mL, 58.7 mmol). After

the slurry had been stirred for 20 min, the solvent was removed under vacuum and the resulting

white residue was extracted with ether (2 × 20 mL). The resulting clear solution was filtered,

and the filtrate was concentrated to ca. 5 mL and placed in a -20 ºC freezer overnight. After 17 h

the colorless needles were isolated and dried for 1 h at 0.1 Torr and 20 ºC. Yield: 0.131 g (57%).

Calc for C8H24NO2B2Na: C, 48.5; H, 10.9; N, 6.28. Found: C, 48.4; H, 10.9; N, 6.32. 1H NMR

(d8-thf): δ 1.25 (1:1:1:1 q, 1JBH = 91 Hz, BH3, 6H), δ 1.35 (pentet, 3JHH = 6.0 Hz, 4-CH2, 2H),

1.63 (m, 3-CH2, 4H), 2.49 (m, 2-CH2, 4H), 3.56 (s, diox, 8H). 13C{1H} NMR (d8-thf): δ 22.7 (s,

4-CH2), 25.5 (s, 3-CH2), 58.9 (s, 2-CH2), 67.8 (s, diox). 11B NMR (d8-thf): δ -14.6 (q, 1JBH = 91

Hz). IR (cm-1): 2355 m, 2396 s br, 2249 s br, 1376 m, 1310 w, 1295 w, 1258 m, 1203 m, 1184

m br, 1158 m br, 1113 m br, 1082 m, 1066 m, 1039 m, 1013 w, 998 w, 972 m, 956 m, 891 m,

872 s, 799 br, 721 br, 669 w, 617 m, 548 w.

Crystallographic Studies.46 Single crystals of each compound were mounted on glass

fibers with Paratone-N oil (Exxon) and immediately cooled to -80 °C in a cold nitrogen gas

stream on the diffractometer. Standard peak search and indexing procedures gave rough cell

dimensions, and least squares refinement yielded the cell dimensions given in Table 4.1. Data

were collected with an area detector by using the measurement parameters listed in Table 4.1.

The measured intensities were reduced to structure factor amplitudes and their estimated

Page 122: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

112

standard deviations by correction for background and polarization effects. Systematically absent

reflections were deleted and symmetry equivalent reflections were averaged to yield the set of

unique data. Unless otherwise specified, all unique data were used in the least squares

refinement. The analytical approximations to the scattering factors were used, and all structure

factors were corrected for both real and imaginary components of anomalous dispersion. No

corrections for decay or secondary extinction were necessary, unless otherwise specified. The

structure was solved by direct methods (SHELXTL). Subsequent least-squares refinement and

difference Fourier calculations revealed the positions of the remaining non-hydrogen atoms. In

the final cycle of least squares, independent anisotropic displacement factors were refined for the

non-hydrogen atoms, unless otherwise specified. Hydrogen atoms were generally observable in

the difference maps. The locations of borane hydrogen atoms were allowed to refine freely. All

other hydrogen atoms were placed in idealized positions; the methyl groups were allowed to

rotate about the C-N axis to find the best least-squares positions. The displacement parameters

for borane and methylene hydrogens were set equal to 1.2 times Ueq for the attached non-

hydrogen atom; those for methyl hydrogens were set to 1.5 times Ueq. Successful convergence

was indicated by the maximum shift/error of 0.000 for the last cycle. Final refinement

parameters are given in Table 4.1. A final analysis of variance between observed and calculated

structure factors showed no apparent errors.

Details unique to each structure are summarized below.

Na[(H3B)2NEt2](diox), 3. Single crystals of Na[(H3B)2NEt2](diox) were grown from thf.

The systematic absences hkl (h + k ≠ 2n) and h0l (l ≠ 2n) were consistent with the space groups

Cc and C2/c. The average values of the normalized structure factors suggested the space group

C2/c, and this choice was confirmed by successful refinement of the proposed model. A face-

Page 123: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

113

indexed absorption correction was applied, the minimum and maximum transmission factors

being 0.953 and 0.980. The quantity minimized by the least-squares program was Σw(Fo2 - Fc

2)2,

where w = {[σ(Fo2)]2 + (0.070P)2}-1 and P = (Fo

2 + 2Fc2)/3. The largest peak in the final Fourier

difference map (0.16 eÅ-3) was located 0.76 Å from H23.

Na[(H3B)2NC5H10](diox)·diox, 4. Single crystals of Na[(H3B)2NC5H10](diox)·diox were

grown from thf. The systematic absences hkl (h + k ≠ 2n) and h0l (l ≠ 2n) were consistent with

the space groups Cc and C2/c. The average values of the normalized structure factors suggested

the space group C2/c, and this choice was confirmed by successful refinement of the proposed

model. A face-indexed absorption correction was applied, the minimum and maximum

transmission factors being 0.975 and 0.993. The quantity minimized by the least-squares

program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 + (0.069P)2 }-1 and P = (Fo

2 + 2Fc2)/3. An

isotropic extinction parameter was refined to a final value of x = 2.81(4) × 10-6 where Fc is

multiplied by the factor k[1 + Fc2xλ3/sin2θ]-1/4 with k being the overall scale factor. The largest

peak in the final Fourier difference map (0.26 eÅ-3) was located 1.25 Å from H3B.

Page 124: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

114

Table 4.1. Crystallographic Data for the New Sodium Aminodiboranate Salts at 193(2) K.

-

Na[(H3B)2NEt2](diox) Na[(H3B)2NC5H10](diox)2

formula C8H24B2NO2Na C13H32B2NO4Na

FW (g mol-1) 210.89 311.01

λ (Å) 0.71073 0.71073

crystal system monoclinic monoclinic

space group C2/c C2/c

a (Å) 13.563(6) 17.805(2)

b (Å) 11.642(5) 10.1123(11)

c (Å) 16.772(7) 20.723(2)

β (deg) 90.691(7) 107.048(7)

V (Å3) 2647.9(19) 3567.2(7)

Z 8 8

ρcalc (g cm-3) 1.058 1.158

μ(mm-1) 0.097 0.101

Rint 0.0763 0.1616

abs correction method face-indexed face-indexed

min., max. transm. factors 0.953, 0.980 0.975, 0.993

data / restraints / params 2525 / 6 / 343 3983 /0 / 209

GOF on F2 0.828 0.906

R1 [I > 2σ(I)]a 0.0397 0.0594

wR2 (all data)b 0.0982 0.16047

max, min Δρelectron (e·Å-3) 0.182/-0.212 0.260/-0.212 aR1 = ∑ | |Fo| - |Fc| | / ∑|Fo| for reflections with Fo

2 > 2 σ(Fo2). bwR2 = [∑w(Fo

2 - Fc2)2 / ∑(Fo

2)2]1/2 for all reflections.

Page 125: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

115

Table 4.2. Selected bond lengths and angles for Na[(H3B)2NEt2](diox), 3.

Bond Lengths (Å)

Na(1)···B(1) 2.926(2) Na(1)-H(12)″ 2.62(2)

Na(1) ··B(2) 2.805(3) Na(1)-H(13)″ 1.16(2)

Na(1) ··B(1)″ 2.682(2) B(1)-N(1) 1.581(2)

Na(1)-O(1) 2.380(1) B(1)-H(11) 1.14(2)

Na(1)-O(2)′ 2.367(1) B(1)-H(12) 1.16(2)

Na(1)-H(11) 2.44(2) B(1)-H(13) 1.16(2)

Na(1)-H(12) 2.62(2) B(2)-N(1) 1.608(2)

Na(1)-H(21) 2.37(2) B(2)-H(21) 1.15(2)

Na(1)-H(22) 2.46(2) B(2)-H(22) 1.12(2)

Na(1)-H(11)″ 2.44(2) B(2)-H(23) 1.10(2)

Bond Angles (deg)

B(1)-Na(1)-B(1)″ 95.65(6) B(1)″-Na(1)-O(2) 123.15(6)

B(1)-Na(1)-B(2) 53.90(6) B(2)-Na(1)-O(1) 95.57(6)

B(1)″-Na(1)-B(2) 148.30(7) B(2)-Na(1)-O(2) 104.49(6)

B(1)-Na(1)-O(1) 131.77(6) O(1)-Na(1)-O(2) 98.30(6)

B(1)″-Na(1)-O(1) 100.67(6) Na(1)-B(1)-Na(1)″ 84.35(6)

B(1)-Na(1)-O(2) 99.89(6) a Symmetry transformations used to generate equivalent atoms: -x+1/2,y+1/2,-z+1/2 and -x+1/2,-y-1/2,-z+1

Page 126: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

116

Table 4.3. Selected bond lengths and angles for Na[(H3B)2NC5H10](diox)·diox, 4.

Bond Lengths (Å)

Na(1)···Na(1)′ 3.844(2) Na(1)-O(1) 2.341(2)

Na(1) ··B(1) 2.872(4) Na(1)-O(2) 2.307(2)

Na(1) ··B(1)′ 2.707(4) B(1)-N(1) 1.576(4)

Na(1) ··B(2) 2.825(4) B(1)-H(11) 1.13(3)

Na(1)-H(11) 2.38(4) B(1)-H(12) 1.04(3)

Na(1)-H(12) 2.41(3) B(1)-H(13) 1.11(3)

Na(1)-H(21) 2.41(3) B(2)-N(1) 1.597(4)

Na(1)-H(22) 2.44(3) B(2)-H(21) 1.20(3)

Na(1)-H(11)′ 2.38(3) B(2)-H(22) 1.21(3)

Na(1)-H(12)′ 2.41(3) B(2)-H(23) 1.21(3)

Na(1)-H(13)′ 2.56(3)

Bond Angles (deg)

B(1)-Na(1)-B(1)′ 93.0(1) O(2)-Na(1)-O(1) 88.20(8)

B(1)-Na(1)-B(2) 53.9(1) O(2)-Na(1)-B(1)′ 101.1(1)

B(1)′-Na(1)-B(2) 144.1(1) O(1)-Na(1)-B(1)′ 102.3(1)

Na(1) -B(1)-Na(1)′ 87.0(1) O(2)-Na(1)-B(2) 101.2(1)

N(1)-B(1)-Na(1)′ 168.2(3) O(1)-Na(1)-B(2) 106.1(1)

N(1)-B(1)-Na(1) 97.6(2) O(2)-Na(1)-B(1) 145.6(1)

O(1)-Na(1)-B(1) 119.3(1) aSymmetry transformations used to generate equivalent atoms: -x+1/2,-y+3/2,-z+ and -x,-y+1,-z+1

Page 127: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

117

References

1. Jensen, J. A.; Gozum, J. E.; Pollina, D. M.; Girolami, G. S. J. Am. Chem. Soc. 1988, 110,

1643-4.

2. Kumar, N.; Yanguas-Gil, A.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. J. Am. Chem. Soc. 2008, 130, 17660-17661.

3. Kumar, N.; Yang, Y.; Noh, W.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2007, 19, 3802-3807.

4. Sung, J.; Goedde, D. M.; Girolami, G. S.; Abelson, J. R. J. Appl. Phys. 2002, 91, 3904-3911.

5. Jayaraman, S.; Yang, Y.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol., A 2005, 23, 1619-1625.

6. Kumar, N.; Noh, W.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2009, 21, 5601-5606.

7. Kumar, N.; Yanguas-Gil, A.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. Appl. Phys. Lett. 2009, 95, 144107/1-144107/3.

8. Yang, Y.; Jayaraman, S.; Sperling, B.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol., A 2007, 25, 200-206.

9. Yang, Y.; Jayaraman, S.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2006, 18, 5088-5096.

10. Yang, Y.; Jayaraman, S.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Cryst. Growth 2006, 294, 389-395.

11. Goedde, D. M.; Girolami, G. S. J. Am. Chem. Soc. 2004, 126, 12230-12231.

Page 128: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

118

12. Jayaraman, S.; Klein, E. J.; Yang, Y.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol. 2005, 23, 631-633.

13. Kim, D. Y.; Yang, Y.; Abelson, J. R.; Girolami, G. S. Inorg. Chem. 2007, 46, 9060-9066.

14. Glass, J. A., Jr.; Kher, S. S.; Spencer, J. T. Chem. Mater. 1992, 4, 530-8.

15. Aldridge, S.; Blake, A. J.; Downs, A. J.; Gould, R. O.; Parsons, S.; Pulham, C. R. J. Chem. Soc., Dalton Trans. 1997, 1007-1012.

16. Bird, P. H.; Wallbridge, M. G. H. J. Chem. Soc. 1965, 3923-8.

17. Haaland, A.; Shorokhov, D. J.; Tutukin, A. V.; Volden, H. V.; Swang, O.; McGrady, G. S.; Kaltsoyannis, N.; Downs, A. J.; Tang, C. Y.; Turner, J. F. C. Inorg. Chem. 2002, 41, 6646-6655.

18. Broach, R. W.; Chuang, I. S.; Marks, T. J.; Williams, J. M. Inorg. Chem. 1983, 22, 1081-4.

19. Borisenko, K. B.; Downs, A. J.; Robertson, H. E.; Rankin, D. W. H.; Tang, C. Y. Dalton Trans. 2004, 967-970.

20. Banks, R. H.; Edelstein, N. M.; Spencer, B.; Templeton, D. H.; Zalkin, A. J. Am. Chem. Soc. 1980, 102, 620-3.

21. Banks, R. H.; Edelstein, N. M.; Rietz, R. R.; Templeton, D. H.; Zalkin, A. J. Am. Chem. Soc. 1978, 100, 1957-8.

22. Giese, H. H.; Habereder, T.; Knizek, J.; Noth, H.; Warchhold, M. Eur. J. Inorg. Chem. 2001, 1195-1205.

23. Filinchuk, Y.; Talyzin, A. V.; Chernyshov, D.; Dmitriev, V. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 76, 092104/1-092104/4.

24. Antsyshkina, A. S.; Sadikov, G. G.; Porai-Koshits, M. A.; Konoplev, V. N.; Sizareva, A. S.; Silina, T. A. Koord. Khim. 1993, 19, 596-600.

Page 129: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

119

25. Babanova, O. A.; Soloninin, A. V.; Stepanov, A. P.; Skripov, A. V.; Filinchuk, Y. J. Phys. Chem. C, 114, 3712-3718.

26. Marynick, D. S.; Lipscomb, W. N. Inorg. Chem. 1972, 11, 820-3.

27. Dain, C. J.; Downs, A. J.; Goode, M. J.; Evans, D. G.; Nicholls, K. T.; Rankin, D. W. H.; Robertson, H. E. J. Chem. Soc., Dalton Trans. 1991, 967-77.

28. Hoekstra, H. R.; Katz, J. J. J. Am. Chem. Soc. 1949, 71, 2488-92.

29. Kim, D. Y.; You, Y.; Girolami, G. S. J. Organomet. Chem. 2008, 693, 981-986.

30. Goedde, D. M.; Windler, G. K.; Girolami, G. S. Inorg. Chem. 2007, 46, 2814-2823.

31. Kim, D. Y.; Girolami, G. S. J. Am. Chem. Soc. 2006, 128, 10969-10977.

32. Beckett, M. A.; Brassington, D. S.; Coles, S. J.; Gelbrich, T.; Hursthouse, M. B. Polyhedron 2003, 22, 1627-1632.

33. David, C. M.; Klein, M. F. J. Chem. Educ. 2001, 78, 952-953.

34. Serrar, C.; Es-Sofi, A.; Boutalib, A.; Ouassas, A.; Jarid, A. Inorg. Chem. 2000, 39, 2224-2226.

35. Titov, L. V.; Gavrilova, L. A.; Zhemchugova, L. V. Russ. Chem. Bull. 1998, 47, 1408-1409.

36. Alcock, N. W.; Burns, I. D.; Claire, K. S.; Hill, A. F. Inorg. Chem. 1992, 31, 4606-10.

37. Deiseroth, H. J.; Sommer, O.; Binder, H.; Wolfer, K.; Frei, B. Z. Anorg. Allg. Chem. 1989, 571, 21-8.

38. Levicheva, M. D.; Titov, L. V.; Psikha, S. B. Zh. Neorg. Khim. 1987, 32, 510-12.

39. Titov, L. V.; Levicheva, M. D.; Psikha, S. B. Zh. Neorg. Khim. 1984, 29, 668-73.

Page 130: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

120

40. Gaines, D. F.; Hildebrandt, S. J. Met. Interact. Boron Clusters 1982, 119-43.

41. Keller, P. C. J. Chem. Soc. D 1969, 1465.

42. Noth, H.; Thomas, S. Eur. J. Inorg. Chem. 1999, 1373-1379.

43. Kim, D. Y. Part 1. Synthesis of metal hydroborates as potential chemical vapor deposition. Part 2. Chemical vapor deposition of titanium-doped magnesium diboraide thin films. Ph.D. Thesis, University of Illinois at Urbana-Champaign, 2007.

44. Daly, S. R. Aminodiboranates: Syntheses, Structural Studies, and Applications for Chemical Vapor Deposition, Ph.D. Thesis, University of Illinois at Urbana-Champaign 2010.

45. Eaton, G. R.; Lipscomb, W. N. N.M.R. Studies of Boron Hydrides and Related Compounds; Benjamin: New York, 1969.

46. Brumaghim, J. L.; Priepot, J. G.; Girolami, G. S. Organometallics 1999, 18, 2139-2144.

Page 131: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

121

Chapter 5: Preparation and Characterization of Alkaline Earth Aminodiboranates

Introduction

Chemical vapor deposition (CVD) and atomic layer deposition (ALD) are excellent

methods to prepare thin films because they can deposit films uniformly on substrates with high

aspect ratios, a feat that is difficult to accomplish by physical vapor deposition (PVD)

techniques.1-11 The composition of the deposited thin films can also be varied depending on the

type of precursors that are used. A prerequisite, however, for the deposition of thin films by

CVD and ALD is that the thin film precursors used must be volatile. Although this requirement

can be achieved for the lighter alkaline earth metals such as magnesium,12-18it is exceedingly

difficult to achieve for the heavier congeners such as barium.19-22

Some transition metal borohydride complexes containing BH4- and B3H8

- ligands are

highly volatile and have proven to be outstanding CVD precursors for the deposition of metal

diboride films.11,23-35 Despite these successes, magnesium and barium complexes of borohydride

ligands such as BH4-, B3H8

-,36-38 and B6H9- 39 do not meet the volatility requirements for CVD,

aside from Mg(B3H8)2(Et2O)2.29 Few volatile barium complexes are known because most

ligands are too small to saturate the coordination sphere of barium and prevent polymeric

bonding modes that greatly decrease vapor pressures. The few known examples of volatile

bariu-m complexes employ large, often multidentate, ligands such as diketonates,40-50

ketoiminates,40-50 alkoxides,30,51,52 cyclopentadienides,53-59 or (pyrazolyl)borates.60-63 In order to

fill remaining voids within the coordination sphere of the barium atom, these ligands are usually

used with neutral ancillary ligands or are modified so that they bear neutral pendant donor

groups.64-66

Page 132: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

122

We have shown that a new class of borohydride ligands known as the aminodiboranates

are extremely versatile ligands for the synthesis of volatile complexes suitable for CVD.67,68 The

ligands can be described as two borohydride groups joined by a amido linker, making them

sterically larger (i.e., able to occupy a larger percentage of a metal’s coordination sphere) than

BH4- and B3H8

-. In particular, volatile complexes of alkaline earths,67,69 d-block,28,69 and f-

block69 metals have been prepared using N,N-dimethylamindiboranate, (H3B)2NMe2-, as a ligand.

Although this ligand forms volatile magnesium complexes, it does not do so with barium. Even

when the barium center is ligated by donors such as diethyl ether, 1,2-dimethoxyethane, or

N,N,N′,N′-tetramethylethylenediamine, the dimethylaminodiboranate complexes are associated in

the solid state69 Lewis base adducts of barium aminodiboranate with more sterically demanding

polydentate ligands such as 12-crown-4 and bis(2-methoxyethyl) ether (diglyme) afford the

monomeric barium complexes Ba[(H3B)2NMe2]2(12-crown-4) and Ba[(H3B)2NMe2]2(diglyme)2,

but these complexes are unfortunately not volatile.69

Here we report the synthesis of several new magnesium aminodiboranate complexes in

which the substituents on nitrogen are varied, and also several barium complexes of the

(H3B)2NMe2- ligand that bear polyamines.

Results and Discussion:

Synthesis of Magnesium and Barium Aminodiboranate Complexes. Grinding MgBr2

and Na[(H3B)2NC5H10] together for 15 min, followed by extraction of the residue with toulene

and distillation of the extract gives a yellow oil. The oil can be sublimed at 0.10 Torr and 50 ºC

to afford the new compound Mg[(H3B)2NC5H10]2, 1, as a white solid

Page 133: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

123

Treatment of a slurry of BaBr2 in a minimal amount of tetrahydrofuran with 2 equiv of

Na[(H3B)2NMe2] or Na[(H3B)2NC5H10] in diethyl ether followed by crystallization from diethyl

ether at -20 ºC yields crystals of the known69 compound Ba[(H3B)2NMe2]2(Et2O)2, 2 or

Ba[(H3B)2NC5H10]2(Et2O)2 7. This procedure is an improvement over the previous route, which

used pure thf instead of a thf/Et2O mixture as the reaction solvent. Drying 2 at room temperature

for an hour at 0.01 Torr gives the partly desolvated analogue Ba[(H3B)2NMe2]2·0.35Et2O, 2′, as

a free flowing white solid. As has been reported previously,69 this latter compound is a useful

starting material for the preparation of other barium dimethylaminodiboranate complexes. Thus,

treatment of 2′ in thf with excess N,N,N′,N″,N″-pentamethylethylenetriamine (pmdta), or

N,N,N′,N′-tetraethylethylenediamine (teed), followed by filtration and cooling to -20 ºC affords

the new complexes Ba[(H3B)2NMe2]2(pmdta), 3, and Ba[(H3B)2NMe2]2(teed), 4, respectively.

The reaction of BaBr2 with Na[(H3B)2NMe2] in tetrahydrofuran, followed by the addition

of 12-crown-4, gives a clear solution from which the unusual salt [Na(12-crown-4)2]-

[Ba((H3B)2NMe2)3(thf)2], 5, can be isolated.

NMR Spectra of Barium Aminodiboranate Complexes. The 1H NMR spectrum of

Ba[(H3B)2NMe2]2·0.35Et2O, 2′, in d8-thf shows two resonances for the aminodiboranate ligand:

a 1:1:1:1 quartet for the borohydride hydrogens at δ 1.70 (1JBH = 93 Hz) and a singlet for the N-

methyl hydrogens at δ 2.24. The 1H NMR spectra of the barium pmdta and teed complexes 3

and 4 in d8-thf show similar features for the aminodiboranate ligands: the 1:1:1:1 quartet for the

BH3 groups appears at δ 1.78 (1JBH = 91 Hz) and δ 1.80 (1JBH = 89 Hz), and the NMe2 singlet

appears at δ 2.25 and 2.24, respectively.

The 1H NMR spectrum of [Na(12-crown-4)2][Ba((H3B)2NMe2)3(thf)2], 5, in d8-thf shows

a 1:1:1:1 quartet for the borohydride hydrogens at δ 1.66 (1JBH = 91 Hz), and a singlet for the N-

Page 134: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

124

methyl hydrogens on the aminodiboranate ligand at δ 2.21. The 13C{1H} NMR spectrum of 5 in

d8-thf shows a resonance at δ 53.3 for the NMe2 group, and characteristic resonances for the thf

and 12-crown-4 ligands. These chemical shifts are not significantly different from those seen for

the electricially neutral complexes above, and so there is no NMR feature that is diagnostic of

the presence of metalate salts containing the aminodiboranate ligand, at least for barium.

The 1H NMR spectrum of the piperidinyl complex Ba[(H3B)2NC5H10]2(thf)2, 6, in d8-thf

shows a broad 1:1:1:1 quartet for the borohydride hydrogens on the aminodiboranate ligand at δ

1.71 (1JBH = 90 Hz), and multiplets at δ 2.50, 1.63, and 1.40 for the 2-, 3-, and 4-methylene

hydrogens of the piperidinyl group, respectively. The 13C{1H}spectrum of 7 in d8-thf exhibits

five resonances at δ 22.5, 25.4, 58.6 due to the 2-, 3-, and 4- carbons of the piperidyl group,

respectively, besides resonances due to thf.

The 1H NMR resonances for the amine and ether ligands in 2′-6 appear in characteristic

locations (see Experimental Section).

The 11B NMR spectra of 2′, 3, 4, 5, and 6 in d8-thf all exhibit a binomial quartet, which

appears at δ -8.07 (1JBH = 93 Hz), δ -7.90 (1JBH = 91 Hz), δ -7.82 (1JBH = 89 Hz), δ -8.67 (1JBH =

91 Hz), and δ -10.8 (1JBH = 90 Hz), respectively. Clearly, there is relatively little variation in

these chemical shifts.

If the structures of these Ba complexes are similar to those seen in the solid state (see

below), then several different environments for the BH3, and NR2 groups should be seen in the

1H NMR spectra; similarly, several of the ligand functional groups should be diastereotopic. In

contrast, the 1H NMR spectra show only one environment for each of these functional groups.

We conclude, not surprisingly, that the complexes must be dynamic on the NMR time scale in

solution.

Page 135: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

125

Crystal Structure of Mg[(H3B)2N(C5H10)]2, 1. Compound 1 (Figure 5.1) crystallizes in

the monoclinic space group P21/n with one molecule in the asymmetric unit. Each

aminodiboranate ligand is bound to the magnesium center by means of four B-H-Mg contacts, so

that the Mg atom has a total coordination number of eight. The eight hydrogen atoms describe a

distorted square antiprism; the planes of the two B-N-B backbones form a dihedral angle of

57.94 (27)° that is half-way between co-planar and orthogonal. This geometry has also been

observed in other aminodiboranate complexes of Mg67,69 and Mn.28 The Mg···H distances fall in

the 1.97-2.09 Å range observed in magnesium complexes of borohydride52,70-72and other

aminodiboranate ligands.28,67,69

X-ray Crystal Structures of Barium Aminodiboranate Complexes. The complex

Ba[(H3B)2NMe2]2(pmdta), 3 (Figure 5.2), crystallizes in the monoclinic space group P21/n with

one molecule in the asymmetric unit. It is a monomer in which both aminodiboranate ligands

coordinate in a κ2H-H3BNMe2BH3-κ2H fashion and the pmdta binds by means of all three of its

nitrogen atoms. Overall, the barium atom is 11 coordinate: eight hydrogen atoms and three

nitrogen atoms define the inner coordination sphere. If each BH3 group is considered to occupy

one coordination site, however, then the geometry about the Ba center can be considered as a

distorted square-capped trigonal prism in which B3 caps the face formed by B1, B2, B4, and N3.

Page 136: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

126

Figure 5.1. ORETP representation of Mg[(H3B)2NC5H10]2, 1. The 35% probability thermal ellipsoids are shown except for hydrogen atoms, which are represented by arbitrarily sized spheres.

Page 137: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

127

Figure 5.2. ORETP representation of Ba[(H3B)2NMe2]2(pmdta), 3. The 35% probability thermal ellipsoids are shown except for hydrogen atoms, which are represented by arbitrarily sized spheres.

Page 138: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

128

Consistent with this view, the Ba1···B1, Ba1-B2, and Ba1-B4 distances are 3.158(7) – 3.173(7)

Å, but the Ba1-B3 distance of 3.262(9) Å is distinctly longer. The Ba-N distances to the pmdta

ligand are 2.884(4) - 2.916(4) Å.

The Ba[(H3B)2NMe2]2(teed) complex 4 (Figure 5.3) crystallizes in the triclinic space

group P1 with one formula unit per asymmetric unit. In actuality, the complex is a dimer in

which the two halves are related by an inversion center. Each barium center is bound to two

κ2H-H3BNMe2BH3-κ2H ligands and to one chelating N,N,N′,N′-tetraethylethylenediamine ligand.

Each Ba center also forms a bond to a hydrogen atom on one of the chelating aminodiboranate

ligands on the other Ba center. Overall, the barium centers again are 11-coordinate, but this time

the inner coordination sphere consists of nine hydrogen atoms and two nitrogen atoms. If each

BH3 unit is considered to occupy one coordination site (and the one bridging hydrogen atom is

ignored) then the structure can be thought of as a capped trigonal prism in which B(1A) caps the

face formed by B(3), B(1), and B(2).

The Ba···B distances to the aminodiboranate ligand not involved in the bridging

interaction are 3.191(6) and 3.173(6) Å. In contrast, the Ba···B distances to the ligand involved

in the bridging interaction are somewhat longer at 3.243(4) and 3.255(5) Å. The Ba···B distance

to the ligand that chelates to the other metal center is even longer at 3.324(4) Å. The average Ba-

H bond length is 3.22 Å, but the Ba-H bond lengths to the κ2 and κ1 bridging hydrogen atoms

differ by 0.27 Å, with the Ba-κ1H interaction having the shortest bond length at 2.582 (5) Å.

The [Na(12-crown-4)2][Ba[(H3B)2NMe2]3(thf)2] complex 5 (Figure 5.6) crystallizes in

the triclinic space group P1 with two formula units per asymmetric unit. The compound is a

complex salt with a 14-coordinate barium anion and a disordered 8-coordinate sodium cation.

The barium anion can best be thought of as a distorted D2d dodecahedron with B(1), B(2), B(3),

Page 139: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

129

and B(4) making up one of the trapezoids and B(5), B(6), O(1), and O(2) making up the other.

All three of the aminiodiboranate ligands are bound to barium in a κ2H-H3BNMe2BH3-κ2H

fashion. The thf ligands are approximately trans from each other with a O(1)-Ba(1)-O(2) bond

angle of 166.31(9) º. All of the Ba(1)-H(X) bond distances fall in the range of 2.77-2.93 Å.

Page 140: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

130

Figure 5.3. ORETP representation of Ba((H3B)2NMe2]2(teed), 4. The 35% probability thermal ellipsoids are shown. Hydrogen atoms attached to carbon have been omitted for clarity.

Page 141: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

131

Figure 5.4. ORETP representation of the [Ba((H3B)2NMe2)3(thf)2]- anion in 5. The 35% probability thermal ellipsoids are shown. The [Na(12-crown-4)2]+ cation and all hydrogen atoms have been omitted for clarity.

Page 142: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

132

Experimental Section

All operations were carried out in vacuum or under argon using standard Schlenk

techniques. All glassware was dried in an oven at 150 °C, assembled hot, and allowed to cool

under vacuum before use. Tetrahydrofuran, diethyl ether, and pentane were distilled under

nitrogen from sodium/benzophenone and degassed with argon immediately before use. The 12-

crown-4 (Avocado Research) was dried over 4 Å sieves (Aldrich). Anhydrous BaBr2 (Strem),

N,N,N′,N″,N″-pentamethyldiethylenetriamine (Aldrich), and N,N,N′,N′-tetraethylethylenediamine

(Aldrich) were used as received. The salts Na[(H3B)2NMe2],73 Na, and Na[(H3B)2NC5H10]

(Chapter 4) were prepared as described elsewhere.

Elemental analyses were performed by the University of Illinois Microanalytical

Laboratory. The IR spectra were recorded on a Nicolet Impact 410 infrared spectrometer as

Nujol mulls between KBr plates. The NMR data were obtained on a Varian Unity Inova-500NB

at 11.75 T (1H), a General Electric GN300WB instrument at 7.00 T(11B and 11B{1H}), a Varian

Unity 400 instrument at 9.4 T (11B and 11B{1H}), or a Varian Unity-500 spectrometer at 11.75 T

(13C{1H}). Chemical shifts are reported in δ units (positive shifts to high frequency) relative to

SiMe4 (1H) or BF3·Et2O (11B).

Bis(piperidinyldiboranato)magnesium, Mg[(H3B)2NC5H10]2, 1. A mixture of MgBr2

(0.300 g, 1.63 mmol) and Na[(H3B)NC5H10] (0.439 g, 3.26 mmol) was gently agitated in the

solid state with the assistance of 20 stainless steel balls. After 15 min, the sticky white solid was

extracted with toluene (3 × 25 mL) to give a clear solution. The extracts were filtered,

combined, and taken to dryness under vacuum to afford an off white solid. The solid was

sublimed at 65 ºC overnight to yield a white solid. Yield: 0.146 g (36%). Anal. Calc for

C10H32N2B4Mg: C, 48.5; H, 13.0; N, 11.3. Found: C, 48.5; H, 13.5; N, 10.8. IR (cm-1): 2531 s

Page 143: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

133

br, 2462 s br, 2394 s br, 2311 s br, 2301 s br, 2265 s br, 2228 s br, 1183 m br, 1166 m br, 1151 m

br, 1092 m br, 1074 mw br, 1057 mw br, 1036 mw br, 1000 m, 949 m, 888 m.

Bis(N,N-dimethylaminodiboranato)bis(diethylether)barium, Ba[(H3B)2NMe2]2-

(Et2O)2, 2. This preparation is an improvement over the one described previously.69 To a

suspension of BaBr2 (1.00 g, 3.37 mmol) and Na[(H3B)2NMe2] (0.638 g, 6.73 mmol) in

tetrahydrofuran (1 mL) was added diethyl ether (30 mL) to give white slurry. As the reaction

proceeded, the mixture became homogenous. After 17 h, the solvent was removed under vacuum

to afford a white residue. The residue was extracted with diethyl ether (3 × 15 mL). The extracts

were filtered, combined, concentrated to ca. 20 mL, and cooled to -20 ˚C to yield white needles

suitable for X-ray diffraction. The spectroscopic data matched those described earlier.69

Bis(N,N-dimethylaminodiboranato)barium 0.35 Diethyletherate, Ba[(H3B)2NMe2]2

·0.35Et2O, 2′. The mother liquor from the above preparation of 3 was taken to dryness to afford

a white solid. Yield: 0.995 g (96%). Anal. Calc for C5.4H27.5N2B4O0.35Ba: C, 21.1, H, 9.04, N,

9.13. Found: C, 20.8; H, 8.75; N, 8.83. 1H NMR (d8-thf): δ 1.70 (1:1:1:1 q, 1JBH = 93 Hz, BH3,

6H), 2.24 (s, NMe2, 6H). 11B NMR (d8-thf): δ -8.07 (q, 1JBH = 93 Hz). IR (cm-1): 2324 s br,

2239 s br, 1217 m br, 1172 m br, 1148 m br, 964 m, 927 w, 907 w, 818 w, 668 w).

Bis(N,N-dimethylaminodiboranato)(N,N,N′,N″,N″-pentamethyldiethylenetriamine)-

barium, Ba[(H3B)2NMe2]2(pmdta), 3. To a solution of Ba[(H3B)2NMe2]2·0.35Et2O (0.502,

1.64 mmol) in tetrahydrofuran (20 mL) was added N,N,N′,N″,N″-pentamethyldiethylenetriamine

(1.12 mL, 5.37 mmol). After being stirred for 17 h, the clear colorless solution was filtered,

concentrated to ca. 10 mL, and cooled to -20 ºC to yield large colorless prisms. Yield: 0.162 g

(20%). Anal. Calc for C13H47N5B4Ba: C, 34.4; H, 10.4; N, 15.4. Found: C, 34.7; H, 10.8; N,

15.6. 1H NMR (d8-thf): δ 1.78 (1:1:1:1 q, 1JBH = 91 Hz, BH3, 12H), 2.18 (s, NMe2, 12H), 2.24

Page 144: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

134

(s, NMe, 3H), 2.25 (s, NMe2, 12H), 2.39 (m, NCH2, 8H). 11B NMR (d8-thf): δ -7.90 (q, 1JBH = 91

Hz). IR (cm-1): 2393 s br, 2300 s br, 2272 s br, 2246 s br, 1352 w, 1310 w, 1297 w, 1246 w,

1216 m br, 1176 m br, 1148 m br, 1104 m br, 1104 w, 1081 w, 1016 m, 967 w, 930 m, 901 w,

789 w, 773 w.

Bis(N,N-dimethylaminodiboranato)(N,N,N′,N′-tetraethylethylenediamine)barium,

Ba[(H3B)2NMe2]2(teed), 4. To a solution of Ba[(H3B)2NMe2]2·0.35Et2O (0.092g, 0.300 mmol)

in tetrahydrofuran (10 mL) was added N,N,N′,N′-tetraethylethylenediamine (0.210 mL, 0.983

mmol). After being stirred for 17 h, the clear colorless solution was filtered, concentrated to ca. 5

mL, and cooled to -20 ºC to afford large colorless blocks. Yield: 0.071 g (48%). 1H NMR (d8-

thf): δ 1.12 (t, 3JHH = 7.0 Hz, β-Et, 12H), 1.80 (1:1:1:1 q, 1JBH = 89 Hz, BH3, 6H), 2.24 (s,

NMe2, 6H), 2.84 (s, NCH2CH2, 4H), 3.38 (q, 3JHH = 7.0 Hz, α-Et, 8H). 11B NMR (d8-thf): δ -

7.82 (q, 1JBH = 89 Hz).

Bis(12-crown-4)sodium Tris(N,N-dimethylaminodiboranato)bis(tetrahydrofuran)-

barate, [Na(12-crown-4)2][Ba((H3B)2NMe2)3(thf)2], 5. To a suspension of BaBr2 (0.171 g,

0.609 mmol) and Na[(H3B)2NMe2]2 (0.115 g, 1.22 mmol) in tetrahydrofuran (1 mL) was added

diethyl ether (30 mL). The white slurry was stirred for 17 h at room temperature, and then was

filtered. The filtrate was taken to dryness to afford a white solid, which was redissolved in thf

(25 mL) and treated with 12-crown-4 (0.099 mL, 0.609 mmol) to give a clear solution. After 20

minutes, the mixture was filtered, and the filtrate was concentrated to ca. 15 mL and cooled to -

20 ºC to afford colorless blocks. Yield: 0.182 g (41%). Anal. Calc for C30H84N3O10NaBa: C,

41.3; H, 9.71; N, 4.82. Found: C, 41.3; H, 9.87; N, 4.78. 1H NMR(d8-thf): δ 1.66 (1:1:1:1 q,

1JBH = 91 Hz, BH3, 18H), 1.76 (m, β-thf, 8H), 2.21 (s, NMe2, 18H), 3.60 (m, α-thf, 8H), 3.65 (s,

12-crown-4, 32H). 13C{1H} NMR (d8-thf): δ 26.4 (s, β-thf), 53.3 (s, NMe2), 68.2 (s, α-thf), 69.4

Page 145: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

135

(br s, 12-crown-4). 11B NMR (d8-thf): δ-8.67 (q, 1JBH = 91 Hz). IR (cm-1): 2391 s br, 2351 s br,

2291 s br, 2249 s br, 1277 m, 1365 m, 1305 m, 1290 m, 1256 m, 1177 s, 1150 s, 1137 s, 1097 s,

1048 s, 1022 s, 919 s, 887 w, 850 m, 795 w, 554 m.

Crystallographic Studies.74 Single crystals of each compound were mounted on glass

fibers with Paratone-N oil (Exxon) and immediately cooled to -80 °C in a cold nitrogen gas

stream on the diffractometer. Standard peak search and indexing procedures gave rough cell

dimensions, and least squares refinement yielded the cell dimensions given in Table 5.1. Data

were collected with an area detector by using the measurement parameters listed in Table 5.1.

The measured intensities were reduced to structure factor amplitudes and their estimated

standard deviations by correction for background and polarization effects. Systematically absent

reflections were deleted and symmetry equivalent reflections were averaged to yield the set of

unique data. Unless otherwise specified, all unique data were used in the least squares

refinement. The analytical approximations to the scattering factors were used, and all structure

factors were corrected for both real and imaginary components of anomalous dispersion. No

corrections for decay or secondary extinction were necessary. The structure was solved by direct

methods (SHELXTL). Subsequent least-squares refinement and difference Fourier calculations

revealed the positions of the remaining non-hydrogen atoms. In the final cycle of least squares,

independent anisotropic displacement factors were refined for the non-hydrogen atoms, unless

otherwise specified. Hydrogen atoms were generally observable in the difference maps. The

locations of borane hydrogen atoms were allowed to refine freely. All other hydrogen atoms

were placed in idealized positions; the methyl groups were allowed to rotate about the C-N axis

to find the best least-squares positions. The displacement parameters for borane and methylene

hydrogens were set equal to 1.2 times Ueq for the attached non-hydrogen atom; those for methyl

Page 146: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

136

hydrogens were set to 1.5 times Ueq. Successful convergence was indicated by the maximum

shift/error of 0.000 for the last cycle. Final refinement parameters are given in Table 5.1. A

final analysis of variance between observed and calculated structure factors showed no apparent

errors.

Details unique to each structure are summarized below.

Mg[(H3B)2N(C5H10)]2, 1. Single crystals of 1 were obtained by sublimation. The

systematic absences h0l (h + l ≠ 2n) and 0k0 (k ≠ 2n) were uniquely consistent with space group

P21/n. An empirical absorption correction (SADABS) was applied, the minimum and maximum

transmission factors being 0.773 and 0.963. The quantity minimized by the least-squares

program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 + (0.053P)2 + 0.07P}-1 and P = (Fo

2 + 2Fc2)/3.

Analysis of the diffraction intensities suggested slight inversion twinning; therefore, the

intensities were calculated from the equation I = xIa + (1-x)Ib, where x is a scale factor that relates

the volumes of the inversion-related twin components. The scale factor refined to a value of

0.313(2). The largest peak in the final Fourier difference map (0.24 eÅ-3) was located 1.18 Å

from C9.

Ba[(H3B)2NMe2]2(pmdta), 3. Single crystals of 3 were grown from diethyl ether. The

systematic absences 0k0 (k ≠ 2n) and h0l (h + l ≠ 2n) were consistent only with the space group

P21/n. No corrections for crystal decay or absorption were necessary, but a face-indexed

absorption correction was applied, the minimum and maximum transmission factors being 0.532

and 0.593. The quantity minimized by the least-squares program was Σw(Fo2 - Fc

2)2, where w =

{[σ(Fo2)]2 + (0.028P)2}-1 and P = (Fo

2 + 2Fc2)/3. The B-H distances involving the bridging

hydrogen atoms were constrained to be equal within ±0.02 Å, similar restraints were applied to

Page 147: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

137

the B-H distances for the terminal hydrogen atoms. The largest peak in the final Fourier

difference map (0.90 eÅ-3) was located 0.89 Å from H31.

Ba[(H3B)2NMe2]2(teed), 4. Single crystals of 4 were grown from diethyl ether. The

triclinic lattice and the average values of the normalized structure factors suggested the space

group P1, which was confirmed by the success of the subsequent refinement of the proposed

model. A face-indexed absorption correction was applied, the minimum and maximum

transmission factors being 0.542 and 0.767. The quantity minimized by the least-squares

program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2}-1 and P = (Fo

2 + 2Fc2)/3. The largest peak in

the final Fourier difference map (2.21 eÅ-3) was located 1.49 Å from H11A.

[Na(12-crown-4)2][Ba((H3B)2NMe2)3(thf)2], 5. Single crystals of 5 were grown from thf.

The triclinic lattice and the average values of the normalized structure factors suggested the

space group P1, which was confirmed by the success of the subsequent refinement. A face-

indexed absorption correction was applied, the minimum and maximum transmission factors

being 0.652 and 0.884.

The structure was solved using direct methods (SHELXTL). Correct positions for the

barium and the non-hydrogen atoms of the aminodiboranate ligands were deduced from an E-

map. Subsequent least-squares refinement and difference Fourier calculations revealed the

positions of the remaining non-hydrogen atoms. All the 12-crown-4 rings are disordered over

two positions. The C-O and C-C distances within these rings were restrained to 1.47 ± 0.01 and

1.52 ± 0.01 Å, respectively. The disordered rings were assigned site occupancy factors (sofs) of

0.5 except for the ring defined by oxygen atoms O45-O48 and the associated carbon atoms. For

these rings a common site occupancy factor was refined so that the sofs for the two disordered

components summed to one; the sof of the major occupancy site refined to 0.645(5). The

Page 148: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

138

quantity minimized by the least-squares program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 +

(0.035P)2}-1 and P = (Fo2 + 2Fc

2)/3. The analytical approximations to the scattering factors were

used, and all structure factors were corrected for both real and imaginary components of

anomalous dispersion. In the final cycle of least squares, independent anisotropic displacement

factors were refined for the non-hydrogen atoms, although light constraints were applied to

ensure similarity of the displacement factors for the disordered atoms. The positions of

hydrogen atoms attached to boron were refined subject to the constraint that B-H = 1.15 Å.

Hydrogen atoms attached to carbon were placed in idealized positions; the methyl groups were

allowed to rotate about the C-N axis to find the best least-squares positions. The displacement

parameters for methylene and boryl hydrogens were set equal to 1.2 times Ueq for the attached

carbon; those for methyl hydrogens were set to 1.5 times Ueq. Successful convergence was

indicated by the maximum shift/error of 0.007 for the last cycle. Final refinement parameters are

given in Table 5.1. The largest peak in the final Fourier difference map (2.19 eÅ-3) was located

0.97 Å from Ba2.

Page 149: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

139

Table 5.1. Crystallographic Data for the New Barium and Magnesium Complexes at 193(2) K.

1 3 4 5

formula C10H32B4N2Mg C13H47B4N5Ba C28H96B8N8Ba2 C30H84B6N3O10NaBa

FW (g mol-1) 247.93 454.12 906.26 872.19

λ (Å) 0.71073 0.71073 0.71073 0.71073

crystal system monoclinic monoclinic triclinic Triclinic

space group P21/n P21/n P1 P1

a (Å) 6.2785(7) 9.544(3) 9.555(3) 12.7000(7)

b (Å) 14.0366(14) 18.799(6) 12.341(4) 21.3688(11)

c (Å) 19.372(3) 14.610(5) 12.448(4) 21.5299(11)

α (deg) 90 90 64.618(4) 60.409(2)

β (deg) 99.327(8) 95.937(6) 81.444(5) 89.888(3)

γ (deg) 90 90 71.243(5) 72.825(3)

V (Å3) 1684.7(3) 2607.2(14) 1255.6(7) 4780.9(4)

Z 4 4 1 4

ρcalc (g cm-3) 0.977 1.157 1.199 1.212

μ (mm-1) 0.087 1.527 1.584 0.886

Rint 0.1316 0.0477 0.1386 0.1408

min. max. transm. factors 0.773, 0.963 0.532, 0.593 0.542, 0.767 0.652, 0.884

data/restraints/params 3213 / 0 / 203 4479/34/253 4783/0/254 21327/2079/1403

GOF on F2 0.996 0.751 0.99 0.901

R1 [I > 2σ(I)]a 0.584 0.0369 0.0376 0.0468

wR2 (all data)b 0.1352 0.0782 0.0805 0.1123

max, min Δρelectron (e·Å-3) 1.035, -0.531 0.901, -0.713 2.214, -0.929 2.188,-1.271 aR1 = ∑ | |Fo| - |Fc| | / ∑|Fo| for reflections with Fo

2 > 2 σ(Fo2). bwR2 = [∑w(Fo

2 - Fc2)2 / ∑(Fo

2)2]1/2 for all reflections.

Page 150: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

140

Table 5.2. Selected bond lengths and angles for Mg[(H3B)2NC5H10]2, 1.

Bond Lengths (Å)

Mg(1) ··B(1) 2.362(5)   Mg(1)-H(21) 1.99(3)

Mg(1) ··B(2) 2.371(4)   Mg(1)-H(22) 2.02(3)

Mg(1) ··B(3) 2.375(4)   Mg(1)-H(31) 1.99(3)

Mg(1) ··B(4) 2.368(5)   Mg(1)-H(32) 1.98(3)

Mg(1)-H(11) 1.95(3)   Mg(1)-H(41) 1.93(3)

Mg(1)-H(12) 2.07(3)   Mg(1)-H(42) 2.04(3)

Bond Angles (deg)

B(1)-Mg(1)-B(2) 65.09(15)   H(12)-Mg(1)-H(21) 95.5(12)

B(1)-Mg(1)-B(3) 150.67(19)   H(12)-Mg(1)-H(22) 71.8(12)

B(1)-Mg(1)-B(4) 119.88(19)   H(12)-Mg(1)-H(31) 136.3(13)

B(2)-Mg(1)-B(3) 127.33(15)   H(12)-Mg(1)-H(32) 164.4(13)

B(2)-Mg(1)-B(4) 148.28(19)   H(21)-Mg(1)-H(22) 53.8(11)

B(3)-Mg(1)-B(4) 65.60(16)   H(21)-Mg(1)-H(31) 102.7(12)

H(11)-Mg(1)-H(12) 55.7(13)   H(21)-Mg(1)-H(32) 87.7(11)

H(11)-Mg(1)-H(21) 72.7(12)   H(22)-Mg(1)-H(31) 148.9(11)

H(11)-Mg(1)-H(22) 97.9(12)   H(22)-Mg(1)-H(32) 98.3(11)

H(11)-Mg(1)-H(31) 92.4(14)   H(31)-Mg(1)-H(32) 56.9(12)

H(11)-Mg(1)-H(32) 139.3(13)

Page 151: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

141

Table 5.3. Selected Bond Lengths and Angles for Ba[(H3B)2NMe2]2(pmdta), 3.

Bond Lengths (Å)

Ba(1) ··B(1) 3.158(7) Ba(1)-H(11) 2.88(4)

Ba(1) ··B(2) 3.173(7) Ba(1)-H(12) 2.68(4)

Ba(1) ··B(3) 3.262(9) Ba(1)-H(21) 2.71(5)

Ba(1) ··B(4) 3.163(8) Ba(1)-H(22) 2.73(4)

Ba(1)-N(3) 2.905(5) Ba(1)-H(31) 2.63(5)

Ba(1)-N(4) 2.916(4) Ba(1)-H(32) 2.77(5)

Ba(1)-N(5) 2.884(4) Ba(1)-H(41) 2.85(5)

Ba(1)-H(42) 2.88(5)

Bond Angles (deg)

B(1)-Ba(1)-B(2) 49.12(18) B(3)-Ba(1)-B(4) 49.1(2)

B(1)-Ba(1)-B(3) 94.1(2) B(3)-Ba(1)-N(3) 89.56(19)

B(1)-Ba(1)-B(4) 139.2(2) B(3)-Ba(1)-N(4) 148.65(18)

B(1)-Ba(1)-N(3) 91.19(17) B(3)-Ba(1)-N(5) 126.52(18)

B(1)-Ba(1)-N(4) 100.44(17) B(4)-Ba(1)-N(3) 103.40(19)

B(1)-Ba(1)-N(5) 127.32(17) B(4)-Ba(1)-N(4) 120.17(18)

B(2)-Ba(1)-B(3) 111.4(2) B(4)-Ba(1)-N(5) 78.94(17)

B(2)-Ba(1)-B(4) 121.1(2) N(3)-Ba(1)-N(4) 62.71(14)

B(2)-Ba(1)-N(3) 134.47(16) N(3)-Ba(1)-N(5) 117.62(14)

B(2)-Ba(1)-N(4) 98.97(16) N(4)-Ba(1)-N(5) 63.36(13)

B(2)-Ba(1)-N(5) 82.45(16)

Page 152: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

142

Table 5.4. Selected Bond Lengths and Angles for Ba[(H3B)2NMe2]2(teed), 4.a

Bond Lengths (Å)

Ba(1) ··B(1) 3.243(4) Ba(1)-H(12) 2.92(4)

Ba(1) ··B(1)′ 3.324(5) Ba(1)-H(13)′ 2.582(5)

Ba(1) ··B(2) 3.255(5) Ba(1)-H(21) 2.88(4)

Ba(1) ··B(3) 3.191(6) Ba(1)-H(22) 2.83(4)

Ba(1) ··B(4) 3.173(6) Ba(1)-H(31) 2.76(4)

Ba(1)-N(3) 2.970(3) Ba(1)-H(33) 2.73(5)

Ba(1)-N(4) 2.947(4) Ba(1)-H(41) 2.78(5)

Ba(1)-H(11) 2.78(4) Ba(1)-H(42) 2.59(4)

Bond Angles (deg)

B(1)-Ba(1)-B(2) 47.17(13) B(3)-Ba(1)-B(4) 48.58(13)

B(1)-Ba(1)-B(3) 88.43(13) B(3)-Ba(1)-N(3) 151.38(11)

B(1)-Ba(1)-B(4) 136.66(13) B(3)-Ba(1)-N(4) 123.31(12)

B(1)-Ba(1)-N(3) 110.12(11) B(3)-Ba(1)-B(1)′ 85.74(15)

B(1)-Ba(1)-N(4) 125.94(12) B(4)-Ba(1)-N(3) 111.75(11)

B(1)-Ba(1)-B(1)′ 76.06(12) B(4)-Ba(1)-N(4) 85.57(14)

B(2)-Ba(1)-B(3) 107.83(12) B(4)-Ba(1)-B(1)′ 101.91(16)

B(2)-Ba(1)-B(4) 131.55(15) N(3)-Ba(1)-N(4) 67.37(9)

B(2)-Ba(1)-N(3) 100.72(11) N(3)-Ba(1)-B(1)′ 78.274(11)

B(2)-Ba(1)-N(4) 79.94(11) N(4)-Ba(1)-B(1)′ 140.11(9)

B(2)-Ba(1)-B(1)′ 119.66(12) a Symmetry transformations used to generate equivalent atoms: -x,-y+1,-z.

Page 153: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

143

Table 5.5. Selected Bond Lengths and Angles for [Na(12-crown-4)2][Ba((H3B)2NMe2)3(thf)2], 5.

Bond Lengths (Å)

Ba(1) ··B(1) 3.255(5) Ba(1)-H(42) 2.7928(5)

Ba(1) ··B(2) 3.258(6) Ba(1)-H(51) 2.8408(6)

Ba(1) ··B(3) 3.260(5) Ba(1)-H(52) 2.7926(6)

Ba(1) ··B(4) 3.237(5) Ba(1)-H(61) 2.8500(6)

Ba(1) ··B(5) 3.255(6) Ba(1)-H(62) 2.9311(6)

Ba(1) ··B(6) 3.291(6) B(1)-N(1) 1.575(6)

Ba(1)-H(11) 2.8292(5) B(2)-N(1) 1.587(6)

Ba(1)-H(12) 2.7949(5) B(3)-N(2) 1.572(6)

Ba(1)-H(21) 2.8151(6) B(4)-N(2) 1.585(6)

Ba(1)-H(22) 2.8708(6) B(5)-N(3) 1.593(6)

Ba(1)-H(31) 2.7969(5) B(6)-N(3) 1.566(6)

Ba(1)-H(32) 2.8093(5) Ba(1)-O(1) 2.871(3)

Ba(1)-H(41) 2.7686(5) Ba(1)-O(2) 2.867(3)

Bond Angles (deg)

B(1)-Ba(1)-B(2) 46.92(13) B(1)-N(1)-B(2) 110.1(3)

B(1)-Ba(1)-B(3) 100.95(14) B(3)-N(2)-B(4) 112.2(3)

B(1)-Ba(1)-B(4) 144.64(14) B(5)-N(3)-B(6) 112.2(4)

B(1)-Ba(1)-B(5) 100.38(15) B(3)-Ba(1)-O(1) 87.39(12)

B(1)-Ba(1)-B(5) 112.54(15) B(3)-Ba(1)-O(2) 79.61(12)

B(1)-Ba(1)-O(1) 115.79(11) O(1)-Ba(1)-O(2) 166.31(9)

B(1)-Ba(1)-O(2) 71.41(12) Symmetry transformations used to generate equivalent atoms: -x+1,-y,-z+1 and -x,-y,-z+2

Page 154: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

144

References.

1. Rossnagel, S. M. IBM J. Res. Dev. 1999, 43, 163-179.

2. Lee, H.-I.; Shimizu, R. J. Vac. Sci. Technol., B 1998, 16, 2528-2531.

3. Gates, S. M. Chem. Rev. 1996, 96, 1519-1532.

4. Hampden-Smith, M. J.; Kodas, T. T. Chem. Vap. Deposition 1995, 1, 8-23.

5. Jones, A. C.; Hitchman, M. L. Chem. Vap. Deposition 2009, 1-36.

6. Crowell, J. E. J. Vac. Sci. Technol. 2003, 21, S88-S95.

7. Leskela, M.; Ritala, M. Angew. Chem., Int. Ed. 2003, 42, 5548-5554.

8. George, S. M. Chem. Rev. 2010, 110, 111-131.

9. Choy, K. L. ECS Trans. 2009, 25, 59-65.

10. Puurunen, R. L. J. Appl. Phys. 2005, 97, 121301/1-121301/52.

11. Yanguas-Gil, A.; Yang, Y.; Kumar, N.; Abelson, J. R. J. Vac. Sci. Technol. 2009, 27, 1235-1243.

12. Carta, G.; El Habra, N.; Crociani, L.; Rossetto, G.; Zanella, P.; Zanella, A.; Paolucci, G.; Barreca, D.; Tondello, E. Chem. Vap. Deposition 2007, 13, 185-189.

13. Matthews, J. S.; Just, O.; Obi-Johnson, B.; Rees, W. S., Jr. Chem. Vap. Deposition 2000, 6, 129-132.

14. Kuchumov, B. M.; Shevtsov, Y. V.; Semyannikov, P. P.; Filatov, E. S.; Igumenov, I. K. ECS Trans. 2009, 25, 927-934.

Page 155: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

145

15. Liu, J. P.; Limb, J.; Lochner, Z.; Yoo, D.; Ryou, J.-H.; Dupuis, R. D. Phys. Status Solidi 2009, 206, 750-753.

16. Fragala, M. E.; Toro, R. G.; Rossi, P.; Dapporto, P.; Malandrino, G. Chem. Mater. 2009, 21, 2062-2069.

17. Hanna, M.; Wang, S.; Redwing, J. M.; Xi, X. X.; Salama, K. Supercond. Sci. Technol. 2009, 22, 015024/1-015024/7.

18. Lautenschlaeger, S.; Sann, J.; Klar, P. J.; Piechotka, M.; Meyer, B. K. Phys. Status Solidi B 2009, 246, 383-386.

19. Shuster, G.; Kreinin, O.; Lakin, E.; Kuzmina, N. P.; Zolotoyabko, E. Thin Solid Films 2010, 518, 4658-4661.

20. Izumi, T. Oyo Butsuri 2010, 79, 14-19.

21. Wojtczak, W. A.; Fleig, P. F.; Hampden-Smith, M. J. Adv. Organomet. Chem. 1996, 40, 215-340.

22. Matthews, J. S.; Rees, W. S., Jr. Adv. Inorg. Chem. 2000, 50, 173-192.

23. Kumar, N.; Noh, W.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2009, 21, 5601-5606.

24. Kumar, N.; Yanguas-Gil, A.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. Appl. Phys. Lett. 2009, 95, 144107/1-144107/3.

25. Kumar, N.; Yanguas-Gil, A.; Daly, S. R.; Girolami, G. S.; Abelson, J. R. J. Am. Chem. Soc. 2008, 130, 17660-17661.

26. Jayaraman, S.; Klein, E. J.; Yang, Y.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol., 2005, 23, 631-633.

27. Jayaraman, S.; Yang, Y.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol., A 2005, 23, 1619-1625.

Page 156: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

146

28. Kim, D. Y. Part 1. Synthesis of metal hydroborates as potential chemical vapor deposition. Part 2. Chemical vapor deposition of titanium-doped magnesium diboraide thin films. Ph.D. Thesis, University of Illinois at Urbana-Champaign, 2007.

29. Kim, D. Y.; Yang, Y.; Abelson, J. R.; Girolami, G. S. Inorg. Chem. 2007, 46, 9060-9066.

30. Nagamatsu, J.; Nakagawa, N.; Muranaka, T.; Zenitani, Y.; Akimitsu, J. Nature 2001, 410, 63-64.

31. Yang, Y.; Jayaraman, S.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2006, 18, 5088-5096.

32. Yang, Y.; Jayaraman, S.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Cryst. Growth 2006, 294, 389-395.

33. Yang, Y.; Jayaraman, S.; Sperling, B.; Kim, D. Y.; Girolami, G. S.; Abelson, J. R. J. Vac. Sci. Technol., A 2007, 25, 200-206.

34. Kumar, N.; Yang, Y.; Noh, W.; Girolami, G. S.; Abelson, J. R. Chem. Mater. 2007, 19, 3802-3807.

35. Sung, J.; Goedde, D. M.; Girolami, G. S.; Abelson, J. R. J. Appl. Phys. 2002, 91, 3904-3911.

36. Hermanek, S.; Plesek, J. Collect. Czech. Chem. Commun. 1966, 31, 177-89.

37. Levicheva, M. D.; Titov, L. V.; Psikha, S. B. Zh. Neorg. Khim. 1987, 32, 510-12.

38. Titov, L. V.; Levicheva, M. D.; Psikha, S. B. Zh. Neorg. Khim. 1984, 29, 668-73.

39. Denton, D. L.; Clayton, W. R.; Mangion, M.; Shore, S. G.; Meyers, E. A. Inorg. Chem. 1976, 15, 541-8.

40. Tammenmaa, M.; Antson, H.; Asplund, M.; Hiltunen, L.; Leskela, M.; Niinisto, L.; Ristolainen, E. J. Cryst. Growth 1987, 84, 151-4.

Page 157: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

147

41. Purdy, A. P.; Berry, A. D.; Holm, R. T.; Fatemi, M.; Gaskill, D. K. Inorg. Chem. 1989, 28, 2799-803.

42. Gardiner, R.; Brown, D. W.; Kirlin, P. S.; Rheingold, A. L. Chem. Mater. 1991, 3, 1053-9.

43. Wills, L. A.; Wessels, B. W.; Richeson, D. S.; Marks, T. J. Appl. Phys. Lett. 1992, 60, 41-3.

44. Sievers, R. E.; Turnipseed, S. B.; Huang, L.; Lagalante, A. F. Coord. Chem. Rev. 1993, 128, 285-91.

45. Belot, J. A.; Neumayer, D. A.; Reedy, C. J.; Studebaker, D. B.; Hinds, B. J.; Stern, C. L.; Marks, T. J. Chem. Mater. 1997, 9, 1638-1648.

46. Tiitta, M.; Niinisto, L. Chem. Vap. Deposition 1997, 3, 167-182.

47. Saanila, V.; Ihanus, J.; Ritala, M.; Leskela, M. Chem. Vap. Deposition 1998, 4, 227-233.

48. Dhote, A. M.; Meier, A. L.; Towner, D. J.; Wessels, B. W.; Ni, J.; Marks, T. J. J. Vac. Sci. Technol., B: Microelectron. Nanometer Struct.--Process., Meas., Phenom. 2005, 23, 1674-1678.

49. Nilsen, O.; Rauwel, E.; Fjellvag, H.; Kjekshus, A. J. Mater. Chem. 2007, 17, 1466-1475.

50. Wersand-Quell, S.; Orsal, G.; Thevenin, P.; Bath, A. Thin Solid Films 2007, 515, 6507-6511.

51. Chi, Y.; Ranjan, S.; Chou, T.-Y.; Liu, C.-S.; Peng, S.-M.; Lee, G.-H. J. Chem. Soc., Dalton Trans. 2001, 2462-2466.

52. Prust, J.; Most, K.; Muller, I.; Alexopoulos, E.; Stasch, A.; Uson, I.; Roesky, H. W. Z. Anorg. Allg. Chem. 2001, 627, 2032-2037.

53. Vehkamaki, M.; Hatanpaa, T.; Hanninen, T.; Ritala, M.; Leskela, M. Electrochem. Solid-State Lett. 1999, 2, 504-506.

Page 158: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

148

54. Ihanus, J.; Haenninen, T.; Hatanpaeae, T.; Aaltonen, T.; Mutikainen, I.; Sajavaara, T.; Keinonen, J.; Ritala, M.; Leskelae, M. Chem. Mater. 2002, 14, 1937-1944.

55. Hatanpaeae, T.; Vehkamaeki, M.; Mutikainen, I.; Kansikas, J.; Ritala, M.; Leskelae, M. Dalton Trans. 2004, 1181-1188.

56. Ihanus, J.; Haenninen, T.; Hatanpaeae, T.; Ritala, M.; Leskelae, M. J. Electrochem. Soc. 2004, 151, H221-H225.

57. Putkonen, M.; Niinisto, L. Top. Organomet. Chem. 2005, 9, 125-145.

58. Hatanpaa, T.; Ritala, M.; Leskela, M. J. Organomet. Chem. 2007, 692, 5256-5262.

59. Vehkamaki, M.; Hatanpaa, T.; Ritala, M.; Leskela, M.; Vayrynen, S.; Rauhala, E. Chem. Vap. Deposition 2007, 13, 239-246.

60. Condorelli, G. G.; Malandrino, G.; Fragala, I. L. Coord. Chem. Rev. 2007, 251, 1931-1950.

61. Malandrino, G.; Lo Nigro, R.; Fragala, I. L. Chem. Vap. Deposition 2007, 13, 651-655.

62. Saly, M. J.; Heeg, M. J.; Winter, C. H. Inorg. Chem. 2009, 48, 5303-5312.

63. Saly, M. J.; Munnik, F.; Baird, R. J.; Winter, C. H. Chem. Mater. 2009, 21, 3742-3744.

64. Schulz, D. L.; Hinds, B. J.; Neumayer, D. A.; Stern, C. L.; Marks, T. J. Chem. Mater. 1993, 5, 1605-17.

65. Neumayer, D. A.; Belot, J. A.; Feezel, R. L.; Reedy, C.; Stern, C. L.; Marks, T. J.; Liable-Sands, L. M.; Rheingold, A. L. Inorg. Chem. 1998, 37, 5625-5633.

66. Studebaker, D. B.; Neumayer, D. A.; Hinds, B. J.; Stern, C. L.; Marks, T. J. Inorg. Chem. 2000, 39, 3148-3157.

67. Kim, D. Y.; Girolami, G. S. Inorg. Chem. 2010, 49, 4942-4948.

Page 159: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

149

68. Daly, S. R.; Kim, D. Y.; Yang, Y.; Abelson, J. R.; Girolami, G. S. J. Am. Chem. Soc. 2010, 132, 2106-2107.

69. Daly, S. R. Aminodiboranates: Syntheses, Structural Studies, and Applications for Chemical Vapor Deposition, Ph.D. Thesis, University of Illinois at Urbana-Champaign 2010.

70. Lobkovskii, E. B.; Titov, L. V.; Psikha, S. B.; Antipin, M. Y.; Struchkov, Y. T. Zh. Strukt. Khim. 1982, 23, 172-4.

71. Bremer, M.; Linti, G.; Noeth, H.; Thomann-Albach, M.; Wagner, G. E. W. J. Z. Anorg. Allg. Chem. 2005, 631, 683-697.

72. Bremer, M.; Nöth, H.; Thomann, M.; Schmidt, M. Chem. Ber. 1995, 128, 455-60.

73. Nöth, H.; Thomas, S. Eur. J. Inorg. Chem. 1999, 1373-1379.

74. Brumaghim, J. L.; Priepot, J. G.; Girolami, G. S. Organometallics 1999, 18, 2139-2144.

Page 160: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

150

Chapter 6: Structures and Properties of Copper Alkene Complexes. Binding of Different Isomers of Cyclododecatriene to Copper Triflate.

Introduction

Copper alkene complexes play a number of important roles in chemistry and biology,

including serving as the active site in the ethylene receptor ETR 1, which plays a key role in fruit

ripening, seed germination, flowering, and other functions. 1-3 In addition, copper alkene

complexes are useful as catalysts for such reactions as conjugate additions to α,β-unsaturated

ketones,4 the aziridination of alkenes,5-7 and the photoconversion of norbornadiene to

quadricyclane.8 Copper alkene complexes are also useful in chiral separations,9 as reagents for

the activation of dioxygen, 10 for their fluorescent properties,11,12 and ferroelectric behavior.13

Finally, copper alkene complexes have been studied as chemical vapor deposition (CVD)

precursors14 and as atomic layer deposition (ALD) precursors15,16 for copper metal. For CVD

and ALD applications, copper complexes of mono- and dialkenes have been investigated, but

there have been no such studies of copper trialkenes. The purpose of the present investigation is

to characterize such compounds, and to determine the relative copper-alkene binding strengths.

The binding strengths are relevant in film deposition chemistries, because they govern not only

the thermal stability of the precursor and consequently its ease of handling, but also the kinetics

of the surface reactions that convert the precursor to the desired copper film.

Copper compounds bearing one ethylene ligand are well known.10,17-26 There are only

two examples of copper complexes that bear more than one ethylene ligand, (C2H4)2CuAlCl4 and

[Cu(C2H4)3][Al{OC(CF3)3}4].27 Neither of these is volatile enough to serve as a CVD precursor.

There are also two known tris-ethylene complexes of silver and gold.28,29

Page 161: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

151

Little is known about copper complexes of polyenes, especially those in which the copper

center is bound to two or more C=C double bonds. Only seven such compounds have been

crystallography characterized: [Cu(cod)2][ClO4] where cod = 1,5-cyclooctadiene,30 Cu2(cod)2X2

where X = Cl, Br, or I,30 Cu(tbc)(OTf), where tbc = 1,2,5,6,9,10-tribenzocyclododeca-1,5,9-

triene-3,7,11-triyne,31 [Cu(ccc-C8H12)(MeOH)][BF4], and [Cu(ccc-C8H12)][Al{OC(CF3)3}4],

where ccc-C8H12 is cis,cis,cis-1,5,9-cyclododecatriene.32 In many copper complexes of

polyenes, each copper center is bound to only one C=C double bond.33-38 In other copper

complexes of polyenes, the number of metal-alkene interactions has not been established.39-42

The alkene 1,5,9-cyclododecatriene, C12H18, exists in four isomeric forms that differ in

the configurations of the double bonds: these four isomers can be designated as ccc, cct, ctt, and

ttt, where c and t refer to cis and trans, respectively. Three of these four trienes (all except the

ccc isomer) are prepared by the catalytic cyclotrimerization of butadiene; typically, “naked”

nickel serves as the catalyst.43,44 Of the four isomers, the ttt isomer is the principal product of the

nickel-based cyclotrimerization catalysis. Cyclododecatrienes have numerous uses:45 for

example, they can be converted into valuable chemicals such as brominated flame retardants46

and ingredients for the perfume industry,47 and they have also been discussed as starting

materials for the preparation of the polyamide precursors laurolactam and dodecanedioc acid.48-50

In 1973, Salomon and Kochi42 described the isolation of adducts of copper(I) triflate with

the cct, ctt, and ttt isomers of cyclododecatriene. The compounds, which were prepared by

treating a benzene solution of Cu(benzene)(OTf) with the appropriate cyclododecatriene isomer,

were characterized by IR, 1H and 13C NMR spectroscopy.42,51 The Cu(C12H18)(OTf) complexes

are white solids that are stable at room temperature in air for days, but which oxidize upon longer

Page 162: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

152

exposure to the atmosphere. Here we characterize their solid state structures, and assess the

copper-alkene binding energies as a function of the cyclododecatriene isomer.

Results and Discussion

X-ray Crystal Structures of Copper Cyclododecatriene Complexes. Crystals of the

cct, ctt, and ttt isomers of Cu(C12H18)(OTf), 1, were prepared by Kochi’s method;

crystallographic data are reported in Table 6.1 and selected bond distances and angles are

presented in Table 6.2. In all three structures, the copper is bound to all three C=C bonds and to

the oxygen of the triflate ligand in a distorted tetrahedral geometry. In the crystal of ttt-1, the

molecule resides on a mirror plane despite the fact that individual molecules lack mirror

symmetry. This means that each site is modeled as a superposition of opposite enantiomers. A

satisfactory disorder model was devised, but the metric parameters for this compound are

somewhat less precise than for the other two isomers. In crystals of cct-1, the triflate group

shows some conformational disorder but both conformers have similar Cu-O bond distances.

The conformations of the C12H18 ligands in the copper complexes are best understood

from inspection of figure 1. For cct-1 (Figure 2), the two cis double bonds and their vicinal

carbon atoms form a tub-like geometry. In ctt-1 (Figure 3), the C12H18 ligand adopts a distorted

syn-fused tub-chair conformation, with the cis double bond being at the end of the tub farthest

from the fusion points. Interestingly, in this complex, one of the two trans double bonds forms

significantly longer Cu-C distances (2.28 Å) than does the other trans double bond (2.19 Å). In

ttt-1 (Figure 4), the C12H18 ligand adopts an all-chair conformation with local three fold

symmetry. This conformation is similar to that seen in the corresponding Ni0 complex Ni(ttt-

C8H12)(PMe3).52 Comparing the average C=C bond length of 1.340(10) Å in ttt-1 with that of

Page 163: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

153

Table 6.1. Crystallographic Data for the Three Cu(C12H18)(OTf) Isomers.a

Cu(cct-C12H18)(OTf) Cu(ctt-C12H18)(OTf) Cu(ttt-C12H18)(OTf)

Formula C13H18CuF3O3S C13H18CuF3O3S C13H18CuF3O3S

mol wt. 374.87 374.87 374.87

space group Pbca P21/c Cmc21

a, Å 8.1718(5) 7.2459(16) 11.345(3)

b, Å 15.3049(9) 13.279(3) 9.915(3)

c, Å 25.0661(16) 15.835(4) 13.767(4)

β, deg 90 99.011(4) 90

V, Å3 3135.0(3) 1505.0(6) 1548.6(7)

T, K 198(2) 193(2) 193(2)

Z 8 4 4

dcalcd, g cm-3 1.589 1.655 1.608

μ, cm-1 1.563 1.628 1.582

size, mm 0.08 × 0.08 × 0.35 0.05 × 0.10 × 0.40 0.09 × 0.30 × 0.56

diffractometer Siemens SMART

Radiation Mo Kα, λ = 0.710 73 Å

monochromator graphite crystal

no. of rflns, total 34869 11663 9332

no. of rflns, unique 2985 2754 1956

Variables 263 191 242

RF (obsd data) 0.0519 0.0334 0.0284

wR2 (all data) 0.1676 0.0640 0.0506 a R1 = Σ ||Fo|-|Fc||/Σ |Fo|, wR2= {Σ[w(Fo

2-Fc2)2]/Σ[w(Fo

2)2]}1/2

Page 164: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

154

Table 6.2. Selected Bond Distances (Å) and Angles (°) for the Three Cu(C12H18)(OTf) Isomers.a

distance/angle Cu(cct-C12H18)(OTf) Cu(ctt-C12H18)(OTf) Cu(ttt-C12H18)(OTf)

Cu(1)-O(1) 2.108(5) 2.080(2) 2.087(11)

Cu(1)-C(1) 2.230(7) 2.156(3) 2.209(9)

Cu(1)-C(2) 2.208(7) 2.151(3) 2.203(7)

Cu(1)-C(5) 2.267(6) 2.313(3) 2.234(8)

Cu(1)-C(6) 2.259(7) 2.258(3) 2.164(14)

Cu(1)-C(9) 2.232(8) 2.197(3) 2.179(10)

Cu(1)-C(10) 2.199(7) 2.190(3) 2.192(17)

C(1)-C(2) 1.314(10) 1.350(5) 1.346(10)

C(5)-C(6) 1.310(9) 1.338(4) 1.338(10)

C(9)-C(10) 1.274(10) 1.350(4) 1.337(10)

O(1)-Cu(1)-C(1) 114.4(3) 117.0(1) 89.1(4)

O(1)-Cu(1)-C(2) 114.1(2) 105.7(1) 105.2(3)

O(1)-Cu(1)-C(5) 105.3(2) 117.1(1) 96.4(8)

O(1)-Cu(1)-C(6) 93.1(3) 90.5(1) 120.4(7)

O(1)-Cu(1)-C(9) 101.8(3) 98.5(1) 107.3(7)

O(1)-Cu(1)-C(10) 131.6(3) 133.3(1) 121.5(9)

Cu(1)-O(1)-S(1) 125.4(5) 120.0(1) 130.5(9)

a When disorder is present, the distance given is that for the major occupancy site.

Page 165: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

155

Figure 6.1. Idealized structures of the free cyclododecatriene ligands.

cct, C2 ctt, Cs ttt, D3

Page 166: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

156

Figure 6.2. ORTEP representation of Cu(cct-C12H18)(OTf) with 30% probability thermal ellipsoids. The hydrogen atoms and the minor disordered component of the triflate group have been omitted for clarity.

Page 167: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

157

Figure 6.3. ORTEP representation of Cu(ctt-C12H18)(OTf) with 30% probability thermal ellipsoids. The hydrogen atoms have been omitted for clarity.

Page 168: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

158

Figure 6.4. ORTEP representation of Cu(ttt-C12H18)(OTf) with 30% probability thermal ellipsoids. The hydrogen atoms and the second minor component of the triflate group have been omitted for clarity.

Page 169: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

159

1.32 Å in the free ligand53 shows that the C=C bond length is essentially the same within

experimental error.

The Cu-O bond lengths to the triflate ligand are 2.108(5) Å in cct-1, 2.080(2) Å in ctt-1,

and 2.087(11) Å in ttt-1. The average Cu-C distances to the C=C double bonds are 2.232 (cct),

2.211 (ctt), and 2.197 Å (ttt); for comparison, the average Cu-C distances to the C=C double

bonds in the all-cis complexes [Cu(ccc-C8H12)(MeOH)][BF4] and [Cu(ccc-

C8H12)][Al{OC(CF3)3}4] are 2.222(2) and 2.193(4) Å respectively.32 There are no clear trends in

these distances.

Solution State Dynamics of Copper Cyclododecatriene Complexes. Variable

temperature 1H and 13C NMR spectra of the free trienes and their copper complexes have been

measured in order to determine whether or not the complexes engage in chemical exchange

processes. For each free triene, the number of different 13C environments seen at room

temperature is as follows: 6 for cct (three methylenes and three methines), 6 for ctt (ditto), and 2

for ttt (one methylene and one methine).54-56 These numbers are consistent with the following

time-averaged point group symmetries of the free trienes: C2 for cct-1 (the two fold rotation axis

bisects the trans double bond), Cs for ctt-1 (the mirror plane bisects the cis double bond), and D3

for ttt-1 (a three fold rotation axis and three perpendicular two folds).

For the copper coordination complexes, the 13C NMR coordination chemical shifts for the

carbon atoms in the C=C double bonds are ca. -6 ppm (e.g. from δ 132 for free ttt-C8H12 to δ 126

for its copper complex).51 The 1H NMR resonances show more subtle behavior, some moving to

higher frequencies and some to lower frequencies upon complexation.

For the cct and ttt trienes, complexation with a copper triflate unit should lower the

symmetry of the bound C12H18 ring relative to free triene by destroying the two-folds. In

Page 170: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

160

contrast, for the ctt triene, complexation with the copper triflate unit should preserve the

symmetry of the triene ring. Thus, if the copper complexes adopt structures similar to those seen

in the solid state (i.e., the triflate group remains bound to the copper center, and the copper center

lies out of the mean ring plane), we expect to see the following numbers of 13C NMR

environments (ignoring the resonance for the triflate group and ignoring low-barrier

conformational torsions that might lower the symmetry even further): 12 for cct-1, 6 for ctt-1,

and 4 for ttt-1. Experimentally, for solutions of the copper complexes dissolved in C7D8, we see

the following numbers of 13C{1H} NMR environments: 12 for cct-1 and 6 for ctt-1, but only 2

for ttt-1. For each compound, the same number of 13C environments is observed both at room

temperature and at –75 °C. Evidently, for ttt-1, the two faces of the ttt ligand are equivalent on

the NMR time scale.

Identical conclusions were drawn from the variable temperature 1H NMR data: for the

cct-1 (figure 5) and ctt-1 (figure 6) the number of environments agrees with the solid state

structure, whereas for ttt-1 (figure 7) the number of environments is half that expected. Again,

for all three compounds the 1H NMR spectra are essentially temperature independent, with some

slight broadening being noted at the lower temperatures for the ttt complex.

To account for the NMR results, there are two principal kinds of explanations, static and

dynamic: (1) the solution and solid state structures of ttt-1 are different, or (2) the solution and

solid state structures are the same but the compound is undergoing a chemical exchange process

in solution that is fast on the NMR time scale.

For the static case, one possibility is that in solution the triflate anion dissociates to give a

Cu(C12H18) cation, in which the copper center sits exactly at the center of the ring, thus restoring

D3 point group symmetry. This possibility can be ruled out, however, from the observation that

Page 171: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

161

ctt-1 is a non electrolyte in nitromethane: it gives an electrical conductivity of 0.0002 ohm-1 cm2

mol-1 vs 103.5 ohm-1 cm2 mol-1 for the 1:1 electrolyte Et4NBr.57 In addition, vapor pressure

osmometry experiments have also concluded that the triflate ligand in ttt-1 is not dissociated in

benzene.42 Another static explanation of the NMR data is that the ttt ring has dissociated

completely from the copper center. This hypothesis is ruled out by the osmometry result, and

also by the finding the 13C NMR chemical shifts of ttt-1 are distinctly different from those seen

for the free triene (see above). For the dynamic case, one possible mechanism, reversible

dissociation of the triene, was ruled out in separate experiments, which showed that exchange

between the bound and free C12H18 trienes is slow on the NMR time scale at -75 °C (see below).

The mechanism most likely to account for the equivalence of the two faces of the ttt

ligand is movement of the triflate ligand between one side of the molecule and the other,

probably by breaking and then re-forming the Cu-O bond. Simultaneously, the copper atom slips

through the center of the ttt-C12H18 ring and emerges on the other side. This motion does not

require breaking of any of the copper-alkene bonds.

For cct-1, there is no evidence of an exchange process involving similar movement of the

copper atom from one face of the C12H18 ring to the other. Unlike the situation for ttt-1, such a

motion would require breaking of the interactions between the copper atom and the two cis C=C

double bonds, as well as torsional motions to flip these bonds from one side of the mean ring

plane to the other. Evidently, such a dynamic process is slow because the copper-alkene

interactions have high dissociation barriers.

The 63Cu NMR spectrum of ctt-1 at room temperature in benzene consists of a broad

signal at δ 212 with a full width at half maximum of 6000 Hz. In contrast, ttt-1 shows a much

sharper resonance at δ 241 with a full width at half maximum of 500 Hz. The narrower line

Page 172: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

162

Figure 6.5. 1H NMR spectrum of Cu(cct-C12H18)(OTf) in d6-benzene.

Page 173: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

163

Figure 6.6. 1H NMR spectrum of Cu(ctt-C12H18)(OTf) in d6-benzene.

Page 174: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

164

Figure 6.7. 1H NMR spectrum of Cu(ttt-C12H18)(OTf) in d6-benzne.

Page 175: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

165

width of this peak reflects a smaller electric field gradient, which in turn is almost

certainly the result of the higher symmetry of the ttt ligand. Consistent with this hypothesis, no

63Cu NMR signal could be observed for cct-1 at 20 ºC in benzene; presumably its line width is

even larger than 6000 Hz. No significant changes in the signals were observed upon cooling the

samples to -80 ºC in toluene.

Equilibrium Binding Studies. Few studies of the binding constants of alkenes to copper

have previously been reported. The formation constants for the binding of alkenes to copper(I)

centers bearing nitrogen donating ligands alkene have been determined from vapor pressure

measurements.58 Copper-alkene bonding becomes stronger as the nitrogen containing ligands

become more basic. A similar study of the stabilities of copper(I) complexes of alkenes and

other small molecules bearing nitrogen donor ligands demonstrated that there is no significant

lengthening of the C=C double bond and essentially no metal-to-ligand π-back bonding.18,59

Finally, the binding ability of alkenes to the Cu(OTf) fragment has been determined by carrying

out competition studies in which an excess of an alkene was added to Cu(OTf)(VS), where VS is

vinyl sulfonate.60 The binding strengths of the three alkenes they tested decreased in the order

ethylene > 1-butene ≥ 1,3-butadiene.

We have carried out a series of competition experiments in order to determine the relative

binding energies of the three cyclododecatriene isomers to copper. For example, a sample of ttt-

1 was mixed with 3 equivalents of cct-C12H18 in C7D8, the mixture allowed to equilibrate, and

then the concentrations of the two starting materials as well as the reaction products cct-1 and ttt-

C12H18 were measured by integrations of the 13C{1H} NMR peaks. Such competition

experiments were conducted for other combinations of copper complexes and free triene, and the

equilibrium constants for the ligand competition experiments were deduced from the measured

Page 176: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

166

relative concentrations. Proof that competition experiments were under equilibrium conditions

was obtained by showing that the same concentrations were obtained from both reaction

directions. A representative 13C{1H} NMR spectrum of one such competition experiment is

shown in figure 8.

We find the following equilibrium constants:

ttt-1 + ctt-C12H18 ctt-1 + ttt-C12H18 Keq = 15 ± 5.

cct-1+ ctt-C12H18 ctt-1 + cct-C12H18 Keq = 10 ± 4.

ttt-1 + cct-C12H18 cct-1 + ttt-C12H18 Keq = 1.5 ± 0.5.

The free energy changes associated with each of these equilibria are -1.6, -1.4, and -0.2 (all ±

0.1) kcal mol-1, respectively. These competition experiments show that the copper prefers to

bond to the ctt isomer of cyclododecatriene. The equilibrium constants, of course, reflect the

free energies of both the complexes as well as the free trienes.

For free 1,5,9-cyclododecatriene, the relative free energies of the various isomers can be

determined from equilibrium experiments, and a convenient way to establish the equilibrium is

through the use of a metathesis catalyst.. In toluene at 25 °C, the equilibrium concentrations

determined in this way are 900(ttt):100(ctt):2(cct).61 Thus, ttt-C12H18 is the most

thermodynamically stable, with the ctt and cct isomers being 1.3 and 3.6 kcal mol-1 higher in

energy, respectively. (The ccc isomer is even less stable, and it is calculated to be 11.5 kcal mol-

1 higher in energy than the ttt isomer.62) These free energies suggest that there is nothing

particularly special about the free energy of the unligated ctt-C12H18; therefore, the preference for

the Cu center to bind to this isomer must relate to the free energy of the copper ctt complex.

Page 177: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

167

Figure 6.8. 13C{1H} NMR spectrum (C7D8, -70 °C) in the methylene region of the equilibrium mixture obtained by adding 3 equivalents of ctt-C12H18 to Cu(ttt-C12H18)(OTf). Key: ○ = ttt-C12H18, □ = ctt-C12H18, ● = Cu(ttt-C12H18)(OTf), ■ = Cu(ctt-C12H18)(OTf).

Page 178: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

168

When the free energies of the unligated trienes are combined with the free energies for

the exchange reactions described above, we can calculate relative free energies for the copper

cyclododecatriene complexes. Relative to ttt-1, the ctt-1 and cct-1 complexes have free energies

of -0.3 and 3.4 kcal mol-1. For comparison, the relative thermodynamic stabilities of the

nickel(0) complexes of 1,5,9-cyclododecatriene have been calculated by a gradient-corrected

DFT method: the Ni ttt isomer is the most stable, and the Ni ctt and Ni cct complexes have free

energies of 4.0 and 5.2 kcal mol-1, respectively.62,63

These data suggest that, metals in trigonal planar coordination environments are most

stable when bonded to the ttt-C12H18, whereas metals in tetrahedral coordination environments

are most stable when bonded to ctt-C12H18, (although the energy of the corresponding ttt

complex is only slightly higher). We can conclude that the bonding between Cu and ctt-C12H18

is especially favorable.

It is relevant to point out here that cct-C12H18 and ttt-C12H18 must distort to a higher

energy conformation in order to bind effectively to the copper center. (That the cct-C12H18 and

ttt-C12H18 are the only two that must distort is related to the conclusion from the NMR studies

that the symmetries of these two rings – but not that of ctt-C12H18 – are lowered upon

complexation with copper triflate.) The necessity to undergo a conformational change

particularly easy to see for cct-C12H18, in which the two cis double bonds lie on opposite sides of

the mean ring plane, as opposed to the conformation seen crystallographically for this isomer, in

which both cis double bonds are on the same side of the mean ring plane. A conformational

change (albeit a more subtle one) is also necessary for ttt-C12H18 to bind to a copper center

located out of the mean ring plane, but in contrast ctt-C12H18 has a conformation very similar to

that seen in its copper complex. Thus, for cct-C12H18 and ttt-C12H18, binding to the metal is

Page 179: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

169

attended by an increase in the conformational energy of the ring; this increase is a penalty that

destabilizes the resulting copper complexes relative to that of ctt-C12H18. Much smaller

conformational changes are necessary for the cyclododecatriene rings to bind to nickel because

the metal center lies in the mean ring plane, and thus the free energies of the nickel complexes

more closely track those of the unligated triene.

Experimental Section

All operations were carried out under argon or vacuum unless otherwise specified.

Solvents were distilled under nitrogen from sodium and benzophenone. The 1,5,9-

cyclododecatriene isomers, copper(I) oxide, and triflic anhydride were obtained from Aldrich

and used as received.

The IR spectra were recorded on a Nicolet Impact 410 FT-IR instrument as Nujol mulls

between KBr plates. Elemental analyses were performed by the University of Illinois

Microanalytical Laboratory. Electrical conductivities were measured on a YSI model 35

conductance meter and calibrated by showing that the measured conductivity of a 0.100 M

sample of Et4NBr in nitromethane (103.5 ohm-1 cm2 mol-1) was in the range expected for a 1:1

electrolyte.57 NMR spectra were recorded on either a Varian Unity-500 spectrometer at 11.75 T

(variable temperature and 13C NMR spectra), a Varian Unity Inova-500NB at 11.75 T (1H NMR

spectra, 2D Spectra Figures 8-14), or a Varian Unity Inova-600 at 14.1 T (equilibrium studies

and 63Cu NMR spectra). NMR chemical shifts are reported in δ units with positive chemical

shifts at higher frequencies relative to TMS or [Cu(MeCN)4][BF4]. 1H and 13C NMR parameters

for the free trienes reported previously54-56are similar to our findings:

Page 180: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

170

For the cct isomer: 1H NMR (C6D6, 25 °C): δ 5.48 (q, JHH = qm 9.5 Hz, 2H, cis), 5.37

(eight line pattern, 9.5 Hz, 2H, trans), 5.28 (qm, JHH = 9.2 Hz, 2H, cis), 2.03 (broad, 8H), 1.92

(m, 4H), 13C{1H} NMR (C6D6, 25 °C): δ 130.9 (s, 2C) 130.8 (s, 2C), 129.2 (s, 2C), 31.4 (s, 2C),

27.8 (s, 2C), 27.7 (s, 2C).

For the ctt isomer: 1H NMR (C6D6, 25 °C): δ 5.30 (m, 2H), 5.18 (m, 2H), 5.02 (m, 2H),

2.02 (m, 8H), 1.95 (m, 4H), 13C{1H} NMR (C6D6, 25 °C): δ 130.9 (s, 2C), 130.8 (s, 2C), 129.2

(s, 2C), 31.4 (s, 2C), 27.8 (s, 2C), 27.7 (s, 2C).

For the ttt isomer: 1H NMR (C6D6, 25 °C): δ 4.93 (s, 6H), 1.99 (s, 12H) 13C{1H} NMR

(C6D6, 25 °C): δ 132.0 (s, 6C), 33.0 (s, 6C).

(Benzene)(trifluoromethanesulfonato)copper(I). To a slurry of red Cu2O (3.62 g, 25.3

mmol) suspended in benzene (100 mL) was added triflic anhydride (10.0 g, 35.4 mmol). The

mixture was heated to reflux for 4 h, during which time the red slurry became a yellow solution;

a small amount of unreacted Cu2O remained. The yellow solution was filtered while still hot,

and as cooled to room temperature a white solid crystallized out. The solid was collected by

filtration, washed with pentane (3 × 15 mL), and dried in vacuum. Yield: 5.33 g (72%).

Microanalytical and IR data matched those reported by Salomon and co-workers.42

(1,5,9-Cyclododecatriene)(trifluoromethanesulfonato)copper(I). The following

procedure is suitable for all three isomers of 1,5,9-cyclododecatriene we studied. To a slurry of

Cu(C6H6)(OTf) (1.50 g, 5.2 mmol) in benzene (15 mL) was added 1,5,9-cyclododecatriene (0.89

g, 5.5 mmol). After 5 min, the colorless solution was diluted with pentane (30 mL), and the

resulting white solid was collected by filtration, washed with pentane (3 × 15 mL), and dried in

vacuum. Yield: 1.72 g (89 %). Microanalytical and IR data matched those reported by Salomon

and co-workers.42

Page 181: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

171

For cct-1: 1H NMR (C6D6 25 °C): δ 5.63 (br, 2H), 5.32 (br, 2H), 4.59 (br, 2H), 1.68 (br,

6H), 1.35 (br, 6H). 1H NMR (C7D8 -70 °C): δ 5.85 (br, 1H), 5.64 (br, 1H), 5.58 (br, 1H), 5.55

(br, 2H), 5.31 (br, 1 H), 1.94 (br, 1H), 1.68 (br, 6H), 1.52 (br, 3H) 1.92 (br, 2H). 13C{1H} NMR

(C6D6, 25 °C): δ 126.9 (two superposed singlets, 2C), 126.4 (s, 1C), 124.3 (s, 1C), 123.6 (s, 1C),

120.8 (s, 1C), 34.4 (s, 1C), 31.4 (s, 1C), 29.0 (s, 1C), 28.2 (s, 1C), 24.3 (s, 1C), 24.1 (s, 1C).

13C{1H} NMR (C7D8, -78 °C): δ 126.0 (s, 1C), 125.6 (s, 1C), 124.0 (s, 1C), 123.7 (s, 1C), 123.1

(s, 1C), 121.1 (q, JC-F = 320 Hz, CF3), 119.1 (s, 1C), 34.0 (s, 1C), 31.1 (s, 1C), 28.4 (s, 1C), 28.0

(s, 1C), 23.4 (s, 1C), 23.4 (s, 1C). The 63Cu NMR signal could not be observed either at room

temperature or -80 °C.

For ctt-1: 1H NMR (C6D6, 25 °C): δ 5.68 (m, 2H), 5.42 (m, 2H), 4.36 (m, 2H), 1.80 (m,

4H), 1.65 (m, 4H), 1.48 (m, 2H), 1.36 (m, 2H). 1H NMR (C7D8, -70 °C): δ 5.69 (m, 2H), 5.36

(m, 2H), 4.31 (m, 2H), 1.79 (br, 4H), 1.71 (br, 4H), 1.52 (br, 2H), 1.34 (br, 2H). 13C{1H} NMR

(C6D6, 25 °C): δ 133.4 (s, 2C), 126.6 (s, 2C), 118.9 (s, 2C), 35.5 (s, 2C), 30.0 (s, 2C), 29.4 (s,

2C). 13C{1H} NMR (C7D8, -78 °C): δ 132.8 (s, 2C), 125.6 (s, 2C), 117.8 (s, 2C), 35.0 (s, 2C),

29.5 (s, 2C), 28.9 (s, 2C), 121.0 (q, JC-F = 319 Hz, CF3). 63Cu NMR (C6D6, 20 °C): δ 202.9 (s,

fwhm = 6000 Hz).

For ttt-1: 1H NMR (C6D6, 25 °C): δ 5.11 (s, 6H), 1.69 (s, 12H). 1H NMR (C7D8, -70

°C): δ 5.17 (s, 6H), 1.78 (br, 6H), 1.68 (br, 6H). 13C{1H} NMR (C6D6, 25 °C): δ 126.6 (s, 6C),

34.9 (s, 6C). 13C{1H} NMR (C7D8, -78 °C): δ 125.8 (s 6C), 121.1 (q, JC-F = 319 Hz, CF3), 34.4

(s, 6C). 63Cu NMR (C6D6, 25 ºC): δ 241.2 (s, fwhm = 500 Hz).

Equilibrium Studies. In small graduated reaction flask, a selected Cu(C12H18)(OTf)

isomer (0.500 g, 1.4 mmol) was combined with 3 equivalents of a different free triene and the

volume diluted to 5.00 mL with toluene-d8. The samples were allowed to equilibrate for 1 hour

Page 182: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

172

at room temperature under Ar, and the relative equilibrium concentrations were established by

13C(1H) NMR spectroscopy at -70 °C. Based on inversion recovery experiments it was found

that a delay time of 3 seconds was sufficient to allow for full relaxation of all carbon

environments.

Crystallographic Studies.64 Single crystals of all three compounds were grown by

cooling saturated solutions in 1:1 toluene:pentane, and were mounted on glass fibers with

Paratone-N oil (Exxon) and immediately cooled to -75 °C in a cold nitrogen gas stream on the

diffractometer. Standard peak search and indexing procedures gave rough cell dimensions, and

least squares refinement yielded the cell dimensions given in Table 6.1. Data were collected

with an area detector by using the measurement parameters listed in Table 6.1. The measured

intensities were reduced to structure factor amplitudes and their esd’s by correction for

background, scan speed, and Lorentz and polarization effects. Corrections for crystal decay were

unnecessary, but all three data sets were corrected for absorption. Systematically absent

reflections were deleted and symmetry equivalent reflections were averaged to yield the set of

unique data. The structures were solved by direct methods (SHELXTL). Subsequent least-

squares refinement and difference Fourier calculations revealed the positions of all the non-

hydrogen atoms. The analytical approximations to the scattering factors were used, and all

structure factors were corrected for both real and imaginary components of anomalous

dispersion. In the final cycle of least squares, independent anisotropic displacement factors were

refined for the non-hydrogen atoms, except where noted below. Hydrogen atoms were included

in calculated positions with C-H = 0.99 and 0.95 Å for methylene and vinylic carbons,

respectively; the displacement parameters for the hydrogen atoms were set equal to 1.2 times

Ueq for the attached carbon.. Successful convergence was indicated by the maximum shift/error

Page 183: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

173

of 0.000 for the last cycle. A final analysis of variance between observed and calculated

structure factors showed no apparent errors. Final refinement parameters are given in Table 6.1.

Further details of the data collection and refinements are given in the separate sections below.

Cu(cct-C12H18)(OTf). The systematic absences 0kl (k≠2n), h0l (l≠2n), and hk0 (h≠2n)

were only consistent with the space group Pbca and this choice was confirmed by successful

refinement of the proposed model. The quantity minimized by the least-squares program was

Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 + (0.087P)2}-1 and P = (Fo

2 + 2Fc2)/3. The triflate group

was disordered over two sites with the site occupancy factors of the major component being

0.753(7). The sulfur and carbon atoms of the disordered triflate molecules were constrained to

idealized tetrahedral geometries with S-C = 1.77 ± 0.01 Å, S-O = 1.45 ± 0.01 Å, and C-F = 1.32

± 0.01 Å. A similarity constraint was imposed on the displacement parameters of overlapping

disordered atoms. The largest peak in the final Fourier difference map (0.39 eÅ-3) was located

1.33 Å from C9 and C10.

Cu(ctt-C12H18)(OTf). The systematic absences 0k0 (k≠2n) and h0l (l≠2n) were only

consistent with the space group P21/c, and this choice was confirmed by successful refinement of

the proposed model. One reflection (2 0 2) was found to be a statistical outlier and was deleted.

The quantity minimized by the least-squares program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 +

(0.017P)2}-1 and P = (Fo2 + 2Fc

2)/3. An isotropic extinction parameter was refined to a final

value of x = 1.1(3) × 10-6 where Fc is multiplied by the factor k[1 + Fc2xλ3/sin2θ]-1/4 with k being

the overall scale factor. The largest peak in the final Fourier difference map (0.29 eÅ-3) was

located 0.86 Å from H6.

Cu(ttt-C12H18)(OTf). The systematic absences hkl (h+k≠2n) and h0l (l≠2n) were

consistent with the space groups Cmc21, Cmcm, and Ama2. The average values of the

Page 184: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

174

normalized structure factors suggested that the space group was non-centrosymmetric, and the

choice of Cmc21 was confirmed by subsequent successful refinement of the proposed model.

One reflection (2 0 0) was partly obscured by the beamstop and was deleted. The quantity

minimized by the least-squares program was Σw(Fo2 - Fc

2)2, where w = {[σ(Fo2)]2 + (0.017P)2}-1

and P = (Fo2 + 2Fc

2)/3. In the final model, all atoms except the copper atom are disordered. The

carbon atoms of the cyclododecatriene ring were disordered over two sites related by the mirror

plane that runs through the copper center; these atoms were assigned site occupancy factors of

exactly 0.5. Chemically equivalent C-C bond distances within the ring were constrained to be

equal within an esd of 0.01 Å. The atoms of the triflate anion were disordered over four sites,

two of these being unique and each of these unique components being disordered across the

mirror plane. The atoms within each unique component were assigned a common site occupancy

factor, the sum of the SOFs adding to 0.5 (except for those atoms lying on the mirror plane,

which were assigned SOFs adding to 1.0). The site occupancy factor for the major component

plus its mirror-related partner refined to 0.562(7). Chemically equivalent C-F, F···F, S-C, S-O,

and O···O distances were constrained to be equal. The displacement parameters for the

disordered oxygen and fluorine atoms were restrained to be near-isotropic. The largest peak in

the final Fourier difference map (0.26 eÅ-3) was located 0.96 Å from C13. The model has the

correct handedness, as shown by the absolute structure parameter of -0.02(2). No larger

supercell was evident in the diffraction record, disorder was still evident even in monoclinic

subgroups, and no sensible model could be devised in the centrosymmetric space group Cmcm.

Page 185: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

175

Figure 6.9. 1H-1H COSY NMR spectrum (d6-benzene, 25 °C) of Cu(ctt-C12H18)(OTf) distinguishing vinyl protons attached to different cis and trans double bonds.

Page 186: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

176

Figure 6.10. 1H-13C HMQC NMR spectrum (d6-benzene, 25 °C) of Cu(ctt-C12H18)(OTf) distinguishing vinyl protons attached to cis and trans double bonds.

Page 187: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

177

Figure 6.11. 1H-1H COSY NMR spectrum (d6-benzene, 25 °C) of Cu(ttt-C12H18)(OTf) distinguishing vinyl protons attached to different cis and trans double bonds.

Page 188: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

178

Figure 6.12. 1H-13C HMQC NMR spectrum (d6-benzene, 25 °C) of Cu(ttt-C12H18)(OTf) distinguishing vinyl protons attached to cis and trans double bonds.

Page 189: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

179

Figure 6.13. 1H-1H COSY NMR spectrum (d6-benzene, 25 °C) of Cu(cct-C12H18)(OTf) distinguishing vinyl protons attached to different cis and trans double bonds.

Page 190: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

180

Figure 6.14. 1H-1H COSY NMR spectrum (d6-benzene, 25 °C) of Cu(cct-C12H18)(OTf) distinguishing vinyl protons attached to different cis and trans double bonds.

Page 191: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

181

References

1. Rodriguez, F. I.; Esch, J. J.; Hall, A. E.; Binder, B. M.; Schaller, G. E.; Bleeckert, A. B. Science 1999, 283, 996-998.

2. Beyer, E. M.; Blomstrom, D. C. In Plant Growth Substances; Skoog, F., Ed.; Springer-Verlag: New York, 1979.

3. Abeles, F. B.; Morgan, P. W.; Selveit, M. E. Ethylene in Plant Biology; 2 ed.; Academic Press: San Diego, 1994.

4. Perez, P. J.; Diaz-Requejo, M. M. In Comprehensive Organometallic Chemistry III; Elsevier: Amsterdam, 2007; Vol. 2.

5. Li, Z.; Quan, R. W.; Jacobsen, E. N. J. Am. Chem. Soc. 1995, 117, 5889-90.

6. Brandt, P.; Soedergren, M. J.; Andersson, P. G.; Norrby, P.-O. J. Am. Chem. Soc. 2000, 122, 8013-8020.

7. Gillespie, K. M.; Sanders, C. J.; O'Shaughnessy, P.; Westmoreland, I.; Thickitt, C. P.; Scott, P. J. Org. Chem. 2002, 67, 3450-3458.

8. Wang, X.-S.; Zhao, H.; Li, Y.-H.; Xiong, R.-G.; You, X.-Z. Top. Catal. 2005, 35, 43-61.

9. Cavallo, L.; Cucciolito, M. E.; De Martino, A.; Giordano, F.; Orabona, I.; Vitagliano, A. Chem. Eur. J. 2000, 6, 1127-1139.

10. Dai, X.; Warren, T. H. Chem. Commun. 2001, 1998-1999.

11. Zhang, J.; Xiong, R.-G.; Chen, X.-T.; Xue, Z.; Peng, S.-M.; You, X.-Z. Organometallics 2002, 21, 235-238.

12. Zhang, J.; Xiong, R.-G.; Zuo, J.-L.; You, X.-Z. Chem. Commun. 2000, 1495-1496.

13. Qu, Z.-R.; Chen, Z.-F.; Zhang, J.; Xiong, R.-G.; Abrahams, B. F.; Xue, Z.-L. Organometallics 2003, 22, 2814-2816.

Page 192: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

182

14. Rickerby, J.; Steinke, J. H. G. Chem. Rev. 2002, 102, 1525-1549.

15. Park, K.-H.; Marshall, W. J. J. Am. Chem. Soc. 2005, 127, 9330-9331.

16. Park, K.-H.; Bradley, A. Z.; Thompson, J. S.; Marshall, W. J. Inorg. Chem. 2006, 45, 8480-8482.

17. Thompson, J. S.; Whitney, J. F. Inorg. Chem. 1984, 23, 2813-19.

18. Thompson, J. S.; Swiatek, R. M. Inorg. Chem. 1985, 24, 110-13.

19. Dias, H. V. R.; Lu, H.-L.; Kim, H.-J.; Polach, S. A.; Goh, T. K. H. H.; Browning, R. G.; Lovely, C. J. Organometallics 2002, 21, 1466-1473.

20. Dias, H. V. R.; Singh, S.; Flores, J. A. Inorg. Chem. 2006, 45, 8859-8861.

21. Dias, H. V. R.; Wu, J. Eur. J. Inorg. Chem. 2008, 2008, 509-522.

22. Straub, B. F.; Eisentrager, F.; Hofmann, P. Chem. Commun. (Cambridge) 1999, 2507-2508.

23. Suenaga, Y.; Wu, L. P.; Kuroda-sowa, T.; Munakata, M.; Maekawa, M. Polyhedron 1996, 16, 67-70.

24. Masuda, H.; Yamamoto, N.; Taga, T.; Machida, K.; Kitagawa, S.; Munakata, M. J. Organomet. Chem. 1987, 322, 121-9.

25. Dai, J.; Yamamoto, M.; Kuroda-Sowa, T.; Maekawa, M.; Suenaga, Y.; Munakata, M. Inorg. Chem. 1997, 36, 2688-2690.

26. Munakata, M.; Kuroda-Sowa, T.; Maekawa, M.; Nakamura, M.; Akiyama, S.; Kitagawa, S. Inorg. Chem. 1994, 33, 1284-91.

27. Sullivan, R. M.; Liu, H.; Smith, D. S.; Hanson, J. C.; Osterhout, D.; Ciraolo, M.; Grey, C. P.; Martin, J. D. J. Am. Chem. Soc. 2003, 125, 11065-11079.

Page 193: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

183

28. Krossing, I.; Reisinger, A. Angew. Chem., Int. Ed. 2003, 42, 5725-5728.

29. Dias, H. V. R.; Fianchini, M.; Cundari Thomas, R.; Campana Charles, F. Angew Chem Int Ed Engl 2008, 47, 556-9.

30. Kok, J. M.; Skelton, B. W.; White, A. H. J. Cluster Sci. 2004, 15, 365-376.

31. Ferrara, J. D.; Tessier-Youngs, C.; Youngs, W. J. Inorg. Chem. 1988, 27, 2201-2.

32. Chernyshova, E. S.; Goddard, R.; Poerschke, K.-R. Organometallics 2007, 26, 4872-4880.

33. Franceschi, F.; Guardigli, M.; Solari, E.; Floriani, C.; Chiesi-Villa, A.; Rizzoli, C. Inorg. Chem. 1997, 36, 4099-4107.

34. Krisyuk, V. V.; Turgambaeva, A. E.; Rhee, S.-W. Polyhedron 2004, 23, 809-813.

35. Chen, T. Y.; Omnes, L.; Vaisserman, J.; Doppelt, P. Inorg. Chim. Acta 2004, 357, 1299-1302.

36. Schmidt, G.; Behrens, U. J. Organomet. Chem. 1995, 503, 101-9.

37. Baenziger, N. C.; Richards, G. F.; Doyle, J. R. Inorg. Chem. 1964, 3, 1529-35.

38. Baenziger, N. C.; Doyle, J. R. U. S. Dep. Commer., Off. Tech. Serv., PB Rep. 1961, 156075, 8 pp.

39. Haight, H. L.; Doyle, J. R.; Baenziger, N. C.; Richard, G. F. Inorg. Chem. 1963, 2, 1301.

40. Paiaro, G.; Netto, N.; Musco, A.; Palumbo, R. Ric. Sci., Rend., Sez. A 1965, 8, 1441-50.

41. Chow, Y. L.; Cheng, X.-e.; Buono-Core, G. E. Organometallics 1987, 6, 1126-1129.

42. Salomon, R. G.; Kochi, J. K. J. Amer. Chem. Soc. 1973, 95, 1889-97.

Page 194: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

184

43. Jolly, P. W.; Wilke, G. In The Organic Chemistry of Nickel, Vol. 2, Organic Synthesis; Academic Press: New York, 1975.

44. Jolly, P. W. In Comprehensive Organometallic Chemistry, Vol. 8; Wilkinson, G., Stone, F. G. A., Abel, E. W., Eds.; Pergamon: New York, 1982.

45. Kosswig, K. Chem.-Ztg. 1972, 96, 373-83.

46. Smith, K.; Liu, C.-H.; El-Hiti, G. A.; Kang, G. S.; Jones, E.; Clement, S. G.; Checquer, A. D.; Howarth, O. W.; Hursthouse, M. B.; Coles, S. J. Org. Biomol. Chem. 2005, 3, 1880-1892.

47. Munro, D., US Patent 6,551,988 (2003).

48. Balbolov, E.; Skumov, M. J. Mol. Catal. A: Chem. 1999, 137, 77-83.

49. Sporka, K.; Hanika, J.; Ruzicka, V. Tech. Chem. (Prague) 1978, 68, 16-20.

50. Tsyskovskii, V. K. Plast. Massy 1969, 18-21.

51. Salomon, R. G.; Kochi, J. K. J. Organometal. Chem. 1974, 64, 135-43.

52. Hoffmann, E. G.; Jolly, P. W.; Küsters, A.; Mynott, R.; Wilke, G. Z. Naturforsch. B 1976, 31, 1712.

53. Immirzi, A.; Allegra, G. Ren. Acc. Naz. Kincei 1967, 43, 338-349.

54. Rawdah, T. N.; El-Faer, M. Z. Tetrahedron Lett. 1996, 37, 4267-4270.

55. Rawdah, T. N.; El-Faer, M. Z. Tetrahedron Lett. 1995, 36, 3381-3384.

56. Anet, F. A. L.; Rawdah, T. N. J. Am. Chem. Soc. 1978, 100, 5003-5007.

57. Geary, W. J. Coord. Chem. Rev. 1971, 7, 81-122.

Page 195: SYNTHESIS, CHARACTERIZATION, AND REACTIVITY OF …

185

58. Munakata, M.; Kitagawa, S.; Kosome, S.; Asahara, A. Inorg. Chem. 1986, 25, 2622-7.

59. Thompson, J. S.; Whitney, J. F. J. Am. Chem. Soc. 1983, 105, 5488-90.

60. Suzuki, T.; Noble, R. D.; Koval, C. A. Inorg. Chem. 1997, 36, 136-140.

61. Thorn-Csanyi, E.; Ruhland, K. Macromol. Chem. and Phys. 1999, 200, 2606-2611.

62. Tobisch, S. Chem. Eur. J. 2003, 9, 1217-1232.

63. Tobisch, S. Adv. Organomet. Chem. 2003, 49, 167-224.

64. Brumaghim, J. L.; Priepot, J. G.; Girolami, G. S. Organometallics 1999, 18, 2139-2144.