30.01.2015 Views

Pdf - Cornell High Energy Synchrotron Source - Cornell University

Pdf - Cornell High Energy Synchrotron Source - Cornell University

Pdf - Cornell High Energy Synchrotron Source - Cornell University

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

On the Cover:<br />

New testing for small size scales<br />

Professor Matthew Miller and his group from the Sibley School<br />

of Mechanical and Aerospace Engineering standing in front of the load<br />

frame / diffractometer system located in Rhodes Hall, <strong>Cornell</strong> <strong>University</strong>.<br />

Back row (left to right): Michael Ross, Kevin McNelis, and Jay Schuren.<br />

Front row (left to right): Jon Acquaviva, Zachary Peeples, Ben Oswald,<br />

and Professor Miller. The system was designed and built for their high<br />

energy diffraction work at CHESS - Complete article on page 37.<br />

Contents<br />

Staff in Focus<br />

John Kopsa<br />

from the Staff<br />

Welcome the new<br />

Chae Un Kim<br />

Facility Upgrades<br />

Chris Conolly<br />

Facility <strong>High</strong>lights<br />

Education & Outreach<br />

Lora Hine<br />

MacCHESS Advances<br />

Marian Szebenyi<br />

2 CHESS Director’s Report<br />

Sol Gruner<br />

3 G-line Director’s Report<br />

Joel Brock<br />

4 MacCHESS Director’s Report<br />

Marian Szebenyi<br />

6 CHESS Assistant Director’s Report<br />

Ernie Fontes<br />

8 User and Student <strong>High</strong>lights<br />

Ernie Fontes and Kathy Dedrick<br />

11 Staff in Focus<br />

Laura Houghton<br />

13 Facility Upgrades, Networking and Computing<br />

Chris Conolly<br />

17 C-line Upgrades and New Capabilities<br />

Ernie Fontes and Ken Finkelstein<br />

19 D-line Upgrades and New Science Capabilities<br />

Ernie Fontes and Detlef Smilgies<br />

22 Education & Outreach<br />

Lora Hine<br />

25 MacCHESS Advances: Microcrystallography & More<br />

Marian Szebenyi<br />

27 Station F3<br />

Ulrich Englich<br />

28 XPaXS: Improved Data Acquisition & Real-Time Analysis at CHESS<br />

Darren Dale<br />

30 CHESS <strong>High</strong>-pressure B1/B2 Update<br />

Zhongwu Wang<br />

31 The <strong>Cornell</strong> <strong>Energy</strong> Recovery Linac:<br />

update on a future source of coherent hard x-rays<br />

Don Bilderback, Bruce Dunham, Georg Hoffstaetter,<br />

Alexander Temnykh, and Sol Gruner<br />

CHESS News Magazine 2009


The <strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong> (CHESS)...<br />

CHESS - <strong>Cornell</strong> <strong>University</strong><br />

Rt 366 & Pine Tree Road<br />

Ithaca, NY 14853<br />

phone: 607-255-7163<br />

fax: 607-255-9001<br />

...is a user-oriented National Facility to provide state-of-the-art synchrotron radiation facilities to the<br />

scientific community.<br />

Supported by grants from the Division of Materials Research of the National Science Foundation, CHESS<br />

encompasses a multifaceted research and development program which is partly in-house and partly<br />

collaborative, with a wide spectrum of experimental groups from Universities, National Laboratories<br />

and Industry.<br />

Research <strong>High</strong>lights<br />

Mechanical Testing<br />

Matt Miller<br />

Citrine Structure<br />

Sol Gruner<br />

Super BioSAXS<br />

Richard Gillilan<br />

Topography<br />

Richard Jones<br />

Nanoparticles<br />

John Hart<br />

37 New Mechanical Testing Methods for Structural<br />

Materials at Small Size Scales<br />

Matthew P. Miller<br />

43 Alteration of Citrine Structure by Hydrostatic Pressure<br />

Sol Gruner, Buz Barstow, Nozomi Ando,<br />

and Chae Un Kim<br />

45 Putting Color into Surface Diffraction<br />

Detlef Smilgies<br />

48 Self-Assembled Nano-Checkered Thin Films Studied<br />

by Reciprocal Space Mapping at CHESS<br />

Sean O’Malley, Peter Bonanno, Keun hyuk Ahn,<br />

Andrei Sirenko, Alex Kazimirov, Soon-Yong Park,<br />

Yoichi Horibe, and Sang-Wook Cheong<br />

50 Structures from Solutions: Biomolecular Small-Angle<br />

Solution Scattering at MacCHESS<br />

Richard Gillilan<br />

54 Development of X-ray Pixel Array Detectors<br />

Sol Gruner<br />

56 Exploring New Physics of Nanoparticle Supercrystals<br />

by <strong>High</strong> Pressure Small Angle X-ray Diffraction<br />

Zhongwu Wang, Ken Finkelstein,<br />

and Detlef Smilgies<br />

58 Topography of Diamonds at CHESS helps Nuclear<br />

Physics Program at JLAB<br />

Richard Jones, Franz Klein, and Ken Finkelstein<br />

60 Nanoparticle-block Copolymer<br />

Ulrich Wiesner, Laura Houghton, and Sol Gruner<br />

61 Heat-bump Measurements at CHESS<br />

A2 Wiggler Beam<br />

Peter Revesz, Alex Kazimirov, Ivan Bazarov,<br />

Jim Savino, Emmett Windisch, and Christopher<br />

MacGahan<br />

63 Advances in X-ray Microfocusing with Monocapillary<br />

Optics at CHESS<br />

Sterling Cornaby, Thomas Szebenyi,<br />

Heung-Soo Lee, and Don Bilderback<br />

67 Watching Carbon Nanotube Forest Growth<br />

using X-rays<br />

Eric Meshot, Mostafa Bedewy, Sameh Tawfick,<br />

K. Anne Juggernauth, Eric Verploegen, Yongyi<br />

Zhang, Michael De Volder, and John Hart<br />

71 Phase Behavior of Water inside Protein Crystals<br />

Chae Un Kim, Buz Barstow, Mark Tate,<br />

and Sol Gruner<br />

Other Information<br />

The <strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong> is<br />

supported by the National Science Foundation and the<br />

National Institutes of Health/National Institute of<br />

General Medical Sciences under grant DMR 0225180.<br />

MacCHESS is supported by NIH 5 P41 RR001646-24.<br />

Editor: Laura Houghton, lab49@cornell.edu<br />

Co-editor: Marian Szebenyi, dms35@cornell.edu<br />

Layout and design:<br />

Laura Houghton - CHESS, <strong>Cornell</strong> <strong>University</strong><br />

Special thank you to David Schuller<br />

CHESS News Magazine 2009 Page 1


CHESS Director’s Report<br />

Sol Gruner<br />

Artist’s rendition of Wilson Lab showing CESR under the present<br />

Upper Alumni Athletic field.<br />

<strong>Synchrotron</strong> radiation (SR) science is evolving rapidly. When<br />

CHESS started in 1979, synchrotron radiation was largely a<br />

niche area populated by a small number of condensed matter<br />

physicists who worked parasitically at high energy physics<br />

facilities. Today, SR is part of the essential infrastructural fabric<br />

of scientists from many disciplines, who would have been<br />

shocked to know that they would be working productively at<br />

large-scale accelerator-based facilities. For example, modern<br />

biological science stands on the two legs of molecular<br />

structural determination and genetic engineering; the former<br />

is totally dependent on SR protein structure determination. In<br />

consequence, many new dedicated SR storage rings are coming<br />

on-line around the world to serve a growing and incredibly<br />

diverse user community. The size of this community is difficult<br />

to estimate, but certainly numbers in the tens of thousands.<br />

CHESS has contributed greatly to the evolution of SR. The<br />

question we ask ourselves daily is: “How can we best continue<br />

to serve the community and to have a unique impact”<br />

Historically, SR science cut its teeth at three laboratories:<br />

<strong>Cornell</strong> (where SR was first systematically studied in the early<br />

1950’s), DESY in Hamburg, Germany, and SLAC in California.<br />

In each case SR started as a parasitic activity to elementary<br />

particle physics accelerators. Within the last few years all three<br />

labs have undergone a transition to primary use of the on-site<br />

accelerators for photon science. These three laboratories have<br />

historically led the way in SR for user applications, in training<br />

accelerator physicists and SR facility scientists who go on to<br />

populate other facilities, and as incubators for new technology.<br />

So, for example, DESY and SLAC are now commissioning hard<br />

X-ray Free Electron Lasers (as is SPring-8 in Japan), whereas<br />

<strong>Cornell</strong> is developing a hard x-ray <strong>Energy</strong> Recovery Linac (ERL)<br />

x-ray source.<br />

<strong>Cornell</strong> has arguably trained more accelerator physics and<br />

beamline scientist leaders than any other laboratory in the<br />

world. Along the way, much of the technology in use at<br />

accelerator-based facilities was developed. Recent NSF reviews<br />

about the future of SR science at <strong>Cornell</strong> have emphatically<br />

pointed out that these three core missions -- serving SR<br />

users, training SR facility scientists and accelerator physicists,<br />

and development of new SR and accelerator technology –<br />

are continuing strengths of <strong>Cornell</strong>’s SR activity and offer a<br />

uniquely useful and important future for Wilson Lab.<br />

In 2006, Wilson <strong>Synchrotron</strong> Laboratory, consisting of CHESS<br />

and the Laboratory for Elementary-Particle Physics (LEPP), was<br />

formed into a single administrative umbrella called the <strong>Cornell</strong><br />

Laboratory for Accelerator-based ScienceS and Education<br />

(CLASSE). Proposals for continued operation of CHESS and ERL<br />

R&D through 2014 have been extremely favorably reviewed;<br />

we are very optimistic that these activities will continue. These<br />

continuing awards will be very positive steps in our hoped<br />

for evolutionary path, namely, to continue to operate the<br />

storage ring for CHESS while completing ERL R&D. Eventually<br />

we’d like to build a full-scale ERL facility that incorporates and<br />

supersedes the existing facility as the world’s first continuousduty,<br />

coherent hard x-ray source.<br />

The Facility <strong>High</strong>lights section of this News Magazine shines<br />

a spotlight on many CLASSE activities. At the same time, user<br />

science is as productive as ever, as highlighted in the Research<br />

<strong>High</strong>lights section of the News Magazine. Enjoy.<br />

Page 2 CHESS News Magazine 2009


G-Line Director’s Report<br />

Joel Brock<br />

G-line has had another banner year, both in terms of the<br />

exciting forefront science that has been accomplished and in<br />

the truly excellent graduate students who have been trained<br />

and performed their research there. For example, the <strong>Cornell</strong><br />

Center for Materials Research’s Interdisciplinary Research<br />

Group studying the growth of thin-films of chemically<br />

complex materials is operating both pulsed laser and<br />

supersonic molecular beam (SMB) deposition systems in the<br />

G3 experimental station. They are using them to perform in<br />

situ, real-time x-ray structural studies of the growth processes<br />

occurring during deposition. These research programs involve<br />

several <strong>Cornell</strong> faculty research groups (faculty, post-docs,<br />

graduate students, and undergrads) and several of the CHESS<br />

staff scientists. The program studying the growth of thin<br />

films of organic semiconductors via SMB deposition provides<br />

an excellent example of the value of both in situ studies and<br />

the immersion of post-docs and graduate students in the<br />

G-line facility. Supervised by Professors George Malarias<br />

(Materials Science), James Engstrom (Chemical Engineering),<br />

Joel Brock (Applied Physics) and Arthur Woll (CHESS/G-line),<br />

post-doctoral research associates Aram Amassian and Sukwon<br />

Hong and graduate students Vladimir Pozdin, Tushar Desai,<br />

and John Ferguson deposited thin films of pentacene onto<br />

SiO 2<br />

a gate dielectric, treated with hexamethyldisilazane<br />

(HMDS) and fluorinated octyltrichlorosilane (FOTS). The<br />

morphology of pristine, as-deposited thin films was<br />

determined during growth by in situ real-time X-ray reflectivity<br />

and was measured again, ex situ, by atomic force microscopy<br />

(AFM) following aging at room temperature in vacuum, in<br />

N 2<br />

atmosphere, and in ambient air. The films deposited on<br />

HMDS and FOTS undergo significant reorganization under<br />

vacuum or in N 2<br />

atmosphere1. The changes observed indicate<br />

a de-wetting behavior. This work highlights the propensity of<br />

small-molecule thin films to undergo significant molecularscale<br />

reorganization at room temperature when kept in<br />

vacuum or in N 2<br />

atmosphere after deposition, and provides<br />

a cautionary note to anyone investigating the behavior of<br />

organic electronic devices and the relationship of ex situ<br />

studies to the initial growth of ultra-thin molecular films.<br />

The SMB deposition system was designed and built over<br />

the past several years by graduate students in the Engstrom<br />

research group. The data collection was a collaboration led<br />

by Drs. Amassian and Hong with intensive involvement by<br />

the graduate students. Arthur Woll’s development of the<br />

analytical tools necessary to extract reliable coverage data<br />

from the time-dependent x-ray reflectivity data was crucial to<br />

the success of this project. Dr. Amassian has recently accepted<br />

a faculty position at the brand new King Abdullah <strong>University</strong> of<br />

Science and Technology (KAUST) but continues to collaborate<br />

with G-line through the KAUST center at <strong>Cornell</strong>.<br />

References<br />

1. A. Amassian, V.A. Pozdin, T.V. Desai, S. Hong, A.R. Woll, J.D.<br />

Ferguson, J.D. Brock, G.G. Malliaras, and J.R. Engstrom;<br />

“Post-deposition Reorganization of Pentacene Films<br />

Deposited on Low-<strong>Energy</strong> Surfaces”, Journal of Materials<br />

Chemistry 19, 5580-5592 (2009)<br />

AFM images (3×3 μm 2 ) obtained ex situ<br />

from ~8 ML pentacene thin films deposited<br />

from a supersonic source (E i<br />

= 2.5 eV) on<br />

(a) bare SiO 2<br />

and on SiO 2<br />

treated with (b)<br />

HMDS and (c) FOTS. The continuous (color<br />

filled) curves were calculated directly from<br />

the AFM images. The (hollow) bars were<br />

calculated from the X-ray measurements.<br />

(a) and (b) were obtained after aging 12<br />

hours in vacuum and 30 days in air, (c) was<br />

obtained after 6 hours in vacuum and 60<br />

days in air. Figure from Ref. [1].<br />

CHESS News Magazine 2009 Page 3


MacCHESS Director’s Report<br />

Marian Szebenyi<br />

“MacCHESS” stands for “Macromolecular<br />

Diffraction at CHESS” (or “son of CHESS”<br />

if one is Scots like Keith Moffat, the<br />

founder of MacCHESS). It was founded<br />

more than 20 years ago, when the<br />

potential of synchrotron sources for<br />

macromolecular structure determination<br />

was just becoming apparent, and was<br />

one of the first organizations dedicated<br />

to welcoming structural biologists into<br />

what had heretofore been a physicists’<br />

domain. Funding was obtained, and has<br />

continued to this day, from the National<br />

Institutes of Health under its National<br />

Center for Research Resources program,<br />

which mandates both service and<br />

R&D activities by supported facilities.<br />

MacCHESS has been among the leaders<br />

in the development and implementation<br />

of the beamline hardware – detectors,<br />

goniometers, cooling systems, cameras<br />

– and software – for alignment, data<br />

collection, processing – that have made<br />

synchrotron work so routine that users<br />

can concentrate on biology and not<br />

details of the structure determination<br />

process. Today, MacCHESS continues<br />

to provide excellent user support, while<br />

its research activities have branched<br />

out into new paths which will enhance<br />

opportunities for structural investigation<br />

beyond the traditional crystallographic<br />

experiment.<br />

Support is provided by MacCHESS for all<br />

biological crystallography at CHESS, i.e.<br />

almost all experiments at stations A1, F1,<br />

and F2, and about half the use of F3, as<br />

well as occasional experiments at other<br />

stations. The typical MacCHESS user<br />

visits for a 24-hour period and collects<br />

5-20 data sets during this time. In the<br />

last year for which statistics are available<br />

(April 2007 – March 2008), there were<br />

approximately 500 badges issued<br />

to MacCHESS users and nearly 100<br />

MacCHESS-related publications released.<br />

Important work by users includes Eddy<br />

Arnold’s ongoing studies of drugs to<br />

treat AIDS 1 , Kate Ferguson’s investigation<br />

of anti-cancer drugs 2 , Gino Cingolani’s<br />

study of how a bacteriophage punctures<br />

a cell membrane 3 , and many more.<br />

Our facilities for crystallographic data<br />

collection are always being improved;<br />

some recent upgrades include:<br />

motorized CCD detector motion,<br />

enhanced crystal centering interface,<br />

private networks at each station for<br />

data transfer, and robust automatic<br />

safety shields. More details are given<br />

in the report on page 25. Mail-in data<br />

collection is available; it has been<br />

used roughly half a dozen times per<br />

year and is working well. A number<br />

of collaborations between MacCHESS<br />

scientists and other investigators are<br />

active and have produced excellent work<br />

such as the elucidation of the active site<br />

structure in the multifunctional enzyme<br />

CD38 4 . Additionally, experiments that<br />

are not standard crystallography can<br />

now be carried out at CHESS:<br />

• Small angle x-ray scattering (SAXS)<br />

and wide-angle x-ray scattering<br />

(WAXS) of solutions, at G1 (or F2)<br />

and F1, respectively. A new sample<br />

cell with a disposable insert, and<br />

temperature control between 4<br />

ºC and 60 ºC, is available. Staff are<br />

knowledgeable in data collection<br />

and processing techniques.<br />

• Pressure cryocooling of crystals<br />

or other samples. This technique<br />

can produce better quality crystals<br />

than those cryocooled at ambient<br />

pressure, and can also stabilize<br />

ligands which are otherwise<br />

disordered. Apparatus for<br />

pressurizing samples with helium,<br />

and cooling them to liquid nitrogen<br />

temperature under pressure, is<br />

installed at CHESS, and trained<br />

staff will process users’ samples, on<br />

request. Pressure-cooling using<br />

other gases is also possible using<br />

equipment in the Gruner lab, located<br />

in nearby Clark Hall.<br />

• Microbeam, produced by focusing<br />

x-rays with a glass capillary.<br />

Capillaries with focal spot sizes of 18<br />

and 5 μm are routinely available, and<br />

custom capillaries can be produced if<br />

needed. Microbeams have been used<br />

to examine small crystals, small good<br />

regions on heterogeneous crystals,<br />

and non-crystalline samples such as<br />

plant fibers.<br />

Research projects at MacCHESS<br />

are focused on 5 initiatives:<br />

Microcrystallography, Pressure<br />

Cryocooling, SAXS and Envelope<br />

Phasing, X-ray Optics (at F3), and<br />

Automation. More detailed reports are<br />

found elsewhere in this Newsmagazine;<br />

here I will just mention a few highlights:<br />

• Dan Schuette (a graduate student<br />

of Sol Gruner) and MacCHESS staff<br />

demonstrated that a prototype Pixel<br />

Array Detector (PAD; one of several<br />

being developed by the Gruner<br />

group) could be used to collect<br />

crystallographic data, of a quality<br />

comparable to that from a CCD, but<br />

with a much shorter readout time.<br />

This PAD is small compared to current<br />

CCD and image plate detectors, but<br />

work is in progress at Area Detector<br />

Systems Corp. to produce a large,<br />

tiled, device for crystallographic use.<br />

See report on detectors on page 54.<br />

Page 4 CHESS News Magazine 2009


• Xinguo Hong, Visiting Scientist, designed and constructed the improved sample cell for SAXS experiments, and also<br />

developed a method for avoiding radiation damage by stepwise translation of the cell between exposures. Moreover, he<br />

has demonstrated the feasibility of combining SAXS and WAXS data to obtain information about the internal structure of<br />

macromolecules.<br />

• Sterling Cornaby (a graduate student of CHESS Associate Director Don Bilderback) collaborated with MacCHESS staff to show<br />

that Laue data collected from many small crystals, using a tailored 30% bandpass beam, could be used to solve structures. See<br />

full report on page 63.<br />

• Chae Un Kim has used the pressure-cryocooling technique, and lots of beamtime, to collect data on samples cooled at various<br />

pressures and examined at various temperatures, to improve our understanding of that most common but most complex<br />

material – water. See page 71.<br />

In a recent competitive renewal process, the MacCHESS proposal received an excellent rating from NIH, and funding is assured<br />

through 2013. MacCHESS personnel tend to stay with the organization for a long time, but there have been some changes<br />

in recent years. Quan Hao, Director since 2001, has moved on to a faculty position at Hong Kong <strong>University</strong>; he continues a<br />

connection with MacCHESS as a collaborator. Marian Szebenyi has taken over as Director. Xinguo Hong has been in residence as<br />

a Visiting Scientist for the last few years and has been a great asset to the MacCHESS SAXS program; he is now at Brookhaven Lab.<br />

Long time technical support person Chris Heaton has retired to the sunny South and been replaced by Scott Smith as the local<br />

CompuMotor expert (among other talents). We welcome Chae Un Kim as a new Staff Scientist; as the principal developer of the<br />

pressure cryocooling technique, he is uniquely qualified to advance research in this area.<br />

References:<br />

1. K. Das, J.D. Bauman, A.D. Clark, Jr., Y.V. Frankel, P.J. Lewi, A.J. Shatkin, S.H. Hughes, and E. Arnold; “<strong>High</strong> Resolution Structures of<br />

HIV-1 Reverse Transcriptase/TMC278 Complexes: Strategic flexibility explains potency against resistance mutations”, PNAS 105,<br />

1466-1471 (2008)<br />

2. J. Schmiedel, A. Blaukat, S. Li, T. Knöchel, and K. M. Ferguson; “Matuzumab Binding to EGFR Prevents the Conformational<br />

Rearrangement Required for Dimerization”, Cancer Cell 13, 365-373 (2008)<br />

3. A. S. Olia, S. Casjens, and G. Cingolani; “Structure of Phage P22 Cell Envelope-penetrating Needle”, Nature Struct. Mol. Biol. 14,<br />

1221-1226 (2007)<br />

4. Q. Liu, I. A. Kriksunov, H. Jiang, R. Graeff, H. Lin, H. C. Lee, and Q. Hao; “Covalent and Noncovalent Intermediates of an NAD<br />

Utilizing Enzyme, CD38”, Chemistry & Biology 15, 1068-1078 (2008)<br />

MacCHESS is pleased to welcome Staff Scientist, Chae Un Kim<br />

Chae Un Kim hails from South Korea,<br />

where he majored in physics at Seoul<br />

National <strong>University</strong> and graduated summa<br />

cum laude in 1999. Following a 2-year<br />

stint in the Korean army, he arrived at<br />

<strong>Cornell</strong> in 2001 as a graduate student in<br />

Sol Gruner’s lab. After completing the<br />

required coursework, he quickly became<br />

interested in proteins under pressure,<br />

an interest of Gruner’s for many years,<br />

but one that involved tricky experiments<br />

whose results were not widely applicable.<br />

Together with fellow grad student<br />

Raphael Kapfer, Chae Un developed a new<br />

apparatus (much easier to use than the<br />

old one) for cryocooling crystals under<br />

pressure, and determined that crystals<br />

cooled in this way, and subsequently<br />

handled at room pressure, were often of<br />

better quality than those cooled in the<br />

normal way. After Raphael’s untimely<br />

death in a bicycle accident, Chae Un<br />

continued as the principal developer of<br />

pressure cryocooling, and proceeded<br />

to elaborate a series of extensions to<br />

the technique. Publications reporting<br />

these developments, beginning with the<br />

initial report in 2005, have generated<br />

great excitement in the crystallographic<br />

community. Chae Un is now recognized<br />

as the number one expert in pressure<br />

cryocooling, and has made numerous<br />

presentations on the topic. In 2005, he<br />

received the Oxford Cryosystems Prize for<br />

his poster, “<strong>High</strong> Pressure<br />

Cooling of Protein Crystals<br />

without Cryoprotectants”,<br />

presented at the Annual<br />

Meeting of the American<br />

Crystallographic<br />

Association. In 2007,<br />

he received the Ph.D<br />

degree in the Field<br />

of Biophysics, with a<br />

thesis entitled “<strong>High</strong><br />

Pressure Cryocooling<br />

for Macromolecular<br />

Crystallography”.<br />

Wishing to continue<br />

working with Gruner, and CHESS, Chae<br />

Un stayed on at <strong>Cornell</strong>, first as a postdoc,<br />

and now as a MacCHESS Staff<br />

Scientist. He is continuing to explore the<br />

ramifications of pressure cooling, and the<br />

advances that it makes possible in protein<br />

crystallography and the understanding of<br />

the physics of proteins, water, and other<br />

materials in cryocooled crystals. Besides<br />

conducting exciting research, which he is<br />

always willing to explain to anyone who<br />

is interested, Chae Un is a very helpful<br />

guy: when people heard<br />

about the pressurecooling<br />

technique, many<br />

of them were eager to<br />

try it; since the method<br />

requires special high<br />

pressure apparatus and<br />

is a little tricky to learn,<br />

Chae Un volunteered to<br />

pressure-cool (many!)<br />

collaborators’ crystals.<br />

When a pressure-cooling<br />

apparatus was installed<br />

at CHESS, it was he who<br />

approved its design and<br />

trained MacCHESS staff in operating it.<br />

MacCHESS is pleased to welcome Chae Un<br />

Kim as its newest Staff Scientist.<br />

CHESS News Magazine 2009 Page 5


CHESS Assistant Director’s Report<br />

Ernie Fontes<br />

Some sort of “rough waters” analogy is<br />

needed to describe activities at CHESS<br />

over the past year. The CHESS staff<br />

kept as even a keel as possible despite<br />

two external forces. First, uncertainty<br />

about science funding affected the<br />

laboratory (as well as researchers across<br />

the United States). Budget constraints<br />

heavily throttled capital spending we<br />

had planned for the 2008-2011 renewal<br />

period. This slowed – but did not<br />

stop – progress to improve x-ray user<br />

facilities, as the reader can learn from<br />

other articles in this magazine. Second,<br />

CHESS was directly impacted by the end<br />

of the <strong>High</strong>-<strong>Energy</strong> Physics (HEP) data<br />

collection program and the start of an<br />

accelerator physics program called CESR-<br />

TA – or CESR Test Accelerator. The CESR-<br />

TA project is a jointly-funded NSF-DOE<br />

project to use the <strong>Cornell</strong> accelerator<br />

(CESR) as a test development model for<br />

the damping rings that will ultimately be<br />

part of the International Linear Collider<br />

(ILC). Alterations to CESR caused some<br />

loss of x-ray running days, though the<br />

hardware improvements resulting<br />

from the project will yield long-term<br />

benefits to x-ray users in terms of better<br />

reliability, better alignment and better<br />

stability of x-ray beams.<br />

The quality of the CHESS facilities and<br />

depth of user support depend critically<br />

on the staff. This past year saw three<br />

longtime staff members depart from<br />

the CHESS team. Jeff White took an<br />

early retirement and was excited<br />

to dedicate himself to upgrading<br />

his own home to be a model of<br />

environment-friendly energy<br />

efficiency. Jeff had worked at<br />

CHESS for over 23 years in many<br />

formative roles, among them<br />

leading the operations team, liaison<br />

to the accelerator staff, developer of<br />

the E-line diagnostic beamline, and<br />

head of the CHESS and ERL safety<br />

committees. Holly Manslank was<br />

a CHESS operator for 4 years and<br />

moved on to a technical position<br />

closer to her new home in Geneva<br />

NY. Holly brought to CHESS a keen<br />

sense of organization and logic that<br />

helped the staff see the benefits of<br />

well-organized and documented<br />

Page 6 CHESS News Magazine 2009<br />

procedures for chemical room use, signal<br />

monitoring, cable mapping and labeling.<br />

And last but certainly not least, Virginia<br />

Bizzell retired after 30 years as CHESS<br />

business manager and staff “memory.”<br />

Virginia served both CHESS Directors,<br />

Batterman and Gruner, and all the CHESS<br />

and MacCHESS staff and users with<br />

sincere respect, nurturing and care over<br />

the entire life of the facility. Her service<br />

is sorely missed and we wish her, Jeff,<br />

and Holly all the best.<br />

The CHESS staff strives to constantly<br />

improve the quality of the x-ray user<br />

facility and user support at CHESS.<br />

Those who visit CHESS periodically have<br />

noticed that the running modes of the<br />

facility have changed dramatically over<br />

the past few years. From its inception,<br />

x-ray users had always collected data<br />

while the HEP program was running<br />

the accelerator and collecting particle<br />

physics data. While in this “parasitic<br />

mode,” the accelerator needed<br />

considerable performance tuning and<br />

the x-ray beam positions and stability<br />

were sometimes compromised. This<br />

running mode changed over the past<br />

few years so that CHESS users now<br />

have a dedicated machine during “x-ray<br />

running.” With dedicated running,<br />

accelerator conditions can be optimized<br />

for the best combination of small source<br />

size, long-term stability, and long beam<br />

lifetime. Our goals are to realize as few<br />

interruptions as possible during x-ray<br />

Fig. 1: Historical data<br />

for hours spent per visit<br />

for macromolecular<br />

crystallography users<br />

and, for all users, the<br />

success rate of receiving<br />

scheduled x-ray beamtime.<br />

For non-crystallographic<br />

users, the hours per visit<br />

has consistently averaged<br />

between 4-5 days, over<br />

three times longer than<br />

for crystallographic data<br />

collection.<br />

data collection. We have seen that when<br />

the machine is well conditioned – a<br />

process we improve upon each day –<br />

interruptions for re-injections should be<br />

limited to once every 8-12 hours or more.<br />

We expect this to improve in the future.<br />

To hasten these improvements, we have<br />

formed a study group of both x-ray and<br />

accelerator experts who will identify and<br />

tackle issues related to beam alignment<br />

and stability.<br />

The reliability and use of CHESS have<br />

changed over time. Figure 1 shows<br />

historical trends for the amount of time<br />

that each visiting research group has<br />

spent collecting x-ray data. Also shown<br />

is the reliability during those visits.<br />

Among the many conclusions that can<br />

be drawn the most obvious is that the<br />

reliability of user beamtime has grown<br />

– at or above 95% - which is similar<br />

to other state-of-the-art synchrotron<br />

facilities across the United States. This<br />

is quite an accomplishment given that<br />

the <strong>Cornell</strong> accelerator was designed to<br />

serve the HEP program and continues<br />

to store both electrons and positrons<br />

simultaneously. The flexible design of<br />

the CESR source affords some benefits,<br />

though, allowing the laboratory to<br />

explore creative accelerator physics<br />

development programs such as CESR-TA<br />

and also test novel undulator designs.<br />

Speaking of novelty, other articles in this<br />

magazine highlight exciting upgrades to<br />

CHESS beamlines and instruments over


the recent past. Updating the x-ray beamline front-ends and optics helps improve x-ray performance and add new capabilities<br />

that users should appreciate. For instance, both the F3 and C1 beamlines have newly-installed 800 millimeter-long white beam<br />

mirrors. These help collimate and/or focus x-ray beams in the vertical direction, while at the same time reducing heat loads and<br />

higher-order harmonic content of downstream monochromator optics. The downstream optics for both stations are still not<br />

ready to reap the full benefits, but this will improve as our budget and resource situation improve over the next year. D-line is<br />

the first-ever CHESS beamline that can now operate without any beryllium windows. Traditionally, beryllium windows serve to<br />

separate downstream x-ray optics environments, typically helium volumes, from the delicate ultra-high vacuum of the accelerator<br />

storage ring. Adding a new vacuum optics box to D-line and removing the windows makes it possible for researchers to use highintensity,<br />

low-energy x-ray beams when needed.<br />

Looking forward to continued funding from the NSF, we have plans for scientific and technical initiatives that will benefit users<br />

in many ways. The next year should show progress on many fronts, including designing and building new high-energy, vertical<br />

focusing optics to increase the x-ray intensity by a factor of forty or so for high-pressure research at the upgraded B2 station. This<br />

will dramatically decrease exposure times for angle-dispersive diamond-anvil high-pressure experiments. We would also like<br />

to roll out horizontal-focusing synthetic multilayer optics at a few stations. Best of all would be to engineer a flexible, tunable<br />

substrate for multilayers that would provide a very rare capability – the ability to adjust and optimize the horizontal focus of a<br />

wide-energy-bandpass x-ray beam. The corresponding x-ray intensity increase would enable much more rapid scanning x-ray<br />

fluorescence imaging at the F3 station, for example. We also expect to see many new applications of fast pixel-array detectors, in<br />

collaboration with developments from the Gruner laboratory.<br />

Sol Gruner - CHESS Director<br />

Sol M. Gruner was raised on a farm in Southern New Jersey. He received his undergraduate physics degree from M.I.T. and his Ph.D.,<br />

also in physics, from Princeton <strong>University</strong>. Following receipt of his Ph.D. in 1977, Sol remained in Princeton as a member of the physics<br />

department faculty. He moved to Ithaca in 1997 as Director of the <strong>Cornell</strong> CHESS facility and as a faculty member in the physics<br />

department and the Laboratory of Applied and Solid State Research (LASSP). He presently holds the post of The John L. Wetherill<br />

Professor of Physics. In addition to his work at CHESS, he has an active, independently funded research group in the physics department.<br />

He is a member of the graduate fields of Physics, Applied Physics, Biophysics, and Materials Science and Engineering.<br />

Sol characterizes himself as working across the continuum between biological and condensed matter physics, with a heavy emphasis<br />

on instrumentation and technique development. Areas of particular interest include the biophysics of lipid membranes, the effects<br />

of pressure on biomacromolecules and assemblies, the physical properties of block copolymers and the development of x-ray<br />

instrumentation and methods. He has been involved in the introduction of liquid crystal methods to the study of membrane lipid phases,<br />

and the discovery of a number of block copolymer phases. His group develops x-ray detectors, and much of the technology for the CCD<br />

detectors now used at synchrotrons throughout the world has come from this effort. The group is presently developing silicon-based<br />

detectors for time-resolved x-ray experiments.<br />

Don Bilderback - CHESS Associate Director<br />

After receiving his PhD in physics (with<br />

R. Collella, Purdue, 1975) Don Bilderback<br />

came to <strong>Cornell</strong> and shortly thereafter<br />

became the first CHESS employee under<br />

Bob Batterman, the first CHESS director.<br />

He has pursued a career in x-ray research,<br />

developing x-ray sources (CHESS and<br />

now the <strong>Energy</strong> Recovery Linac), x-ray<br />

detectors (first real-time back-reflection<br />

Laue camera when in Materials Science<br />

and Engineering, see www.multiwire.<br />

com), the first transmission x-ray mirror<br />

and the first bendable mirror with table<br />

legs underneath that you push apart to<br />

accomplish the bending. In 1993, he was a<br />

co-winner of an R&D 100 Award, an honor<br />

given annually to the one hundred most<br />

significant technical innovations of the<br />

year, in recognition of the development<br />

of tapered capillaries for producing<br />

micron-sized x-ray beams. The capillary<br />

work is ongoing and last year resulted in<br />

the excellent thesis by Sterling Cornaby<br />

(Applied and Engineering Physics PhD,<br />

2008) entitled “The Handbook of X-ray<br />

Single-bounce Monocapillary Optics”.<br />

Don has also been concerned with highheat-load<br />

x-ray optics over the last 20<br />

years. His idea of cryo-cooling the silicon<br />

optics has been widely adopted by other<br />

synchrotron laboratories and resulted<br />

in the Compton Award of 1998 to Don<br />

Bilderback, Andreas Freund, Gordon<br />

Knapp and Dennis Mills. Don recently<br />

set the goal (from the backdrop of ERL<br />

development) of making x-ray optics reach<br />

a one nanometer hard x-ray beam size.<br />

The APS/NSLS II community has picked up<br />

this goal and is rapidly developing Laue<br />

lenses that might be one of the first kinds<br />

of optical component that might reach this<br />

goal. This is very exciting!<br />

Don is a member of the ACA, the AAAS,<br />

and is a fellow of the American Physical<br />

Society. He often serves as a consultant to<br />

APS, ESRF and other synchrotron sources.<br />

He is a member of the Photon Science<br />

Committee for the Deutsches Elektronen-<br />

<strong>Synchrotron</strong> (DESY) and of the Nanoprobe<br />

Beam-line Advisory Committee at the<br />

NSLS II project. And Don is currently<br />

the <strong>Cornell</strong> <strong>University</strong> correspondent for<br />

<strong>Synchrotron</strong> Radiation News and a Coeditor<br />

for World Publishing Co. for a book<br />

series on <strong>Synchrotron</strong> Radiation.<br />

Don Bilderback is married to Becky, his<br />

wife and sweetheart of 39 years. He<br />

has one son, Doug, and a 2½ year-old<br />

grandson, Ian, who live in Seattle. Don<br />

loves to travel, commute to work on his<br />

bicycle, swim, and cross-country ski. He<br />

recently was involved (with many other<br />

Ithacans) in settling 53 refuges from<br />

Burma in the Ithaca area. One young lady<br />

from the group, Wimber Pha, lived with<br />

the Bilderbacks five days a week, where<br />

she received tutoring in English, math,<br />

driving a car, etc and moved up to English<br />

as a Second Language in her studies. Don<br />

also enjoys discussing science and faith<br />

issues.<br />

CHESS News Magazine 2009 Page 7


User and Student <strong>High</strong>lights<br />

Ernie Fontes and Kathy Dedrick<br />

Although it may be impossible to<br />

capture the excitement and breadth<br />

of research done by visitors to<br />

CHESS, it is fun trying! With a bias<br />

towards the most recent half year<br />

or so, this brief article will show<br />

some of the rich flavor of work done<br />

by established scientists as well as<br />

aspiring undergraduates. CHESS<br />

“shows off” the successes of our<br />

users in a number of ways, using this<br />

magazine (of course), short highlights<br />

posted on the CHESS web site,<br />

editorial coverage on official <strong>Cornell</strong><br />

news outlets, and in international<br />

arenas using highlights posted at<br />

www.lightsources.org. We frequently<br />

stream highlights to the National<br />

Science Foundation, and some of the<br />

more colorful stories get posted on<br />

their web site and news channels.<br />

Please keep us informed about your<br />

successes so that we can to spread<br />

information to our wide CHESS<br />

community of users and supporters.<br />

This past January Debashis Ghosh’s<br />

lab at the Hauptman-Woodward<br />

Medical Research Institute<br />

(HWI) in Buffalo,<br />

New York<br />

published the three-dimensional structure of aromatase, the key enzyme required for<br />

the body to make estrogen 1 . This structure plus those of two other enzymes involved<br />

in controlling estrogen levels provide the first means to visualize the mechanism of<br />

synthesizing estrogen; the enzymes also offer targets for drugs to treat some types<br />

of breast cancer. Ghosh added 2 ; “Now that we know the structures of all three key<br />

enzymes implicated in estrogen-dependant breast cancers, our goal is to have a<br />

personalized cocktail of inhibitors customized to the specific treatment needs of each<br />

patient. Our knowledge about these three enzymes will enable us to develop three<br />

mutually exclusive inhibitors customized to each patient’s needs which will work in<br />

harmony together with minimal side effects.”<br />

Media channels lit up February 17th with a stunning image (fig. 1) and story about a<br />

three year project by Rice <strong>University</strong> scientist Jane Tao and collaborators, who mapped<br />

out the coat of the PsV-F virus, one of very few viruses containing double-stranded<br />

RNA to be studied at the atomic level. Many diffraction images were needed to create<br />

this detailed picture, which appeared online Feb. 25 in the journal Proceedings of the<br />

National Academy of Sciences 3 . The packing of molecules in the PsV-F coat proved to<br />

be surprisingly different from that in related viruses, a finding that will ultimately help<br />

researchers to understand the structure and function of viruses and uncover ways<br />

to fight viral infections 4-6 . Rice <strong>University</strong> press said: “If a picture is worth a thousand<br />

words, then Rice <strong>University</strong>’s precise new image of a virus’ protective coat is seriously<br />

undervalued. More than three years in the making, the image... could help scientists<br />

find better ways to both fight viral infections and design new gene therapies. The<br />

stunning image…was painstakingly created from hundreds of high-energy X-ray<br />

diffraction images and paints the clearest picture yet of the viruses’ genome-encasing<br />

shell called a ‘capsid’.”<br />

Changing channels to metallurgy and mechanical engineering, Matt Miller (<strong>Cornell</strong>)<br />

has been selected to receive the ASM Henry Marion Howe Medal for 2009 for<br />

his publication 7 , “Measuring Stress Distributions in Ti-6 Al-4V Using<br />

<strong>Synchrotron</strong> X-ray Diffraction.” This award honors the author<br />

whose paper has been selected as the best of those<br />

published in the journal Metallurgical and Materials<br />

Transactions for the prior year. Miller, who reports on<br />

his group’s many efforts in the cover article of this<br />

newsmagazine, points out that the paper cited for<br />

the award covers some of the first titanium data<br />

taken at CHESS and was the basis of the third<br />

chapter of graduate student Joel Bernier’s Ph.D.<br />

thesis. Now that work is responsible for him<br />

having to rent a tuxedo!<br />

Fig. 1: <strong>High</strong>-energy x-ray diffraction was<br />

used to reveal the structure of the protective<br />

protein coat of PsV-F.<br />

Credit: J. Pan & Y.J. Tao/Rice <strong>University</strong> 1<br />

Page 8 CHESS News Magazine 2009


The end of August saw news channels<br />

saturated with stories about a discovery<br />

of a “hidden treasure” in the art world.<br />

Jennifer Mass, chemist and senior scientist<br />

at Delaware’s Winterthur Museum and<br />

Country Estate, gave a keynote presentation<br />

on x-ray fluorescence to study buried<br />

painting layers at the August 19th, 2009<br />

American Chemical Society (ACS) meeting<br />

symposium, “Practical Applications of<br />

Surface Chemistry: Art and Sensing<br />

Applications.” Her presentation recounted<br />

work enabled by a collaboration with CHESS<br />

scientists Arthur Woll, Don Bilderback and<br />

Sol Gruner to develop a confocal x-ray<br />

fluorescence (CXRF) microscope to examine<br />

the chemistry – and ultimately see the<br />

color palette – of layers of paint buried<br />

beneath a later portrait. X-radiography<br />

of an N. C. Wyeth family portrait (c.<br />

1922-1924) revealed the presence of a<br />

second painting buried underneath the<br />

surface. CXRF was able to identify the<br />

chemistry of the buried paint layers and,<br />

ultimately, help Mass reconstruct the color<br />

palette. Mass added: “to date, there have<br />

been no published studies on N.C. Wyeth’s<br />

painting materials and methods, and so this<br />

research will not only elucidate a buried<br />

painting but also contribute to the extant<br />

N.C. Wyeth art historical scholarship and<br />

attribution decisions. The <strong>Cornell</strong> <strong>University</strong><br />

synchrotron source is currently the only<br />

facility for the joint confocal XRFand XRF<br />

intensity mapping of buried paintings.” This<br />

story received streaming video coverage of<br />

the ACS press meeting and was the topic<br />

of the NPR Science Friday radio interview<br />

as well as print coverage by the <strong>Cornell</strong><br />

Chronicle, CHESS and lightsources.org 8-12 .<br />

This summer saw a particularly large (and<br />

welcomed!) group of undergraduate<br />

student research interns at Wilson<br />

Laboratory. CHESS and MacCHESS hosted 7<br />

students in research projects, who filled out<br />

an unusually large summer community of 41<br />

undergraduates including 11 participating<br />

in the CLASSE Research Experiences for<br />

Undergraduates (REU) program, 18 students<br />

in a CMS program in high-energy physics<br />

and 6 students providing technical support<br />

and upgrades for the CESR-TA program.<br />

One REU student, Elizabeth Brost, worked<br />

with scientists Peter Revesz and Don Hartill<br />

on “Vibration Studies at CHESS.” Liza, a<br />

Physics major at Grinnell College, configured<br />

a test station using a seismic accelerometer<br />

with fast-fourier-transform analysis to<br />

characterize mechanical vibrations and<br />

component stability at a number of critical locations – e.g. x-ray optics and<br />

beam position monitors - around the lab. Peter Melich and Adam Putzer<br />

formed a “dynamic duo” working with Arthur Woll to upgrade the confocal x-ray<br />

fluorescence scanning stage to perform “topographic” measurements. Peter and<br />

Adam, both <strong>Cornell</strong> juniors majoring in Applied and Engineering Physics, wrote<br />

Labview programs and machined and configured translation stages, encoders,<br />

and automated dial indicators to create topographic-like scans of 3-dimensional<br />

objects, similar to skimming across a surface as an atomic-force microscope does.<br />

Their hope is to dramatically speed up x-ray fluorescence scans of surfaces by<br />

following the interfacial surface directly, rather than having to scan an entire 3D<br />

volume (fig. 2).<br />

Fig. 2: Rapid, “topographic-like” x-ray fluorescence imaging involves following the<br />

surface (below) instead of having to scan the full 3D volume of the specimen (top).<br />

Michael Lyons, a <strong>Cornell</strong> junior majoring in Electrical and Computer Engineering,<br />

worked with Phil Sorensen, Bob Seeley and Ernie Fontes to help design and<br />

build a computer-automated apparatus that will be used to carefully control<br />

the heating and cooling process of vacuum equipment. The so-called “bakeout”<br />

process is used to heat vacuum equipment to high temperatures to accelerate<br />

the removal of adsorbed or trapped water from vacuum vessels and parts. Mike<br />

made progress using the EPICS control system to automate heating and reading<br />

temperatures and pressures. Paul Grigas, a junior studying Operations Research<br />

at <strong>Cornell</strong>, worked with Laura Houghton, Kathy Dedrick, Phil Sorensen and Ernie<br />

Fontes to examine ways to improve web-enabled databases to track the user<br />

community at CHESS. CHESS uses databases to keep track of the hundreds<br />

of student and scientist visitors each year who share dozens of scientific<br />

instruments. Paul explored modern software options and identified three<br />

candidate packages that are being considered for a next generation system.<br />

Gavrielle Untracht, a <strong>Cornell</strong> junior majoring in Applied and Engineering<br />

Physics, worked with Tom Szebenyi and Don Bilderback to improve the<br />

fabrication of tapered-glass capillary x-ray optics. She learned how to run the<br />

computerized glass puller that makes x-ray optics for CHESS use (the optics<br />

make micron diameter x-ray beams for our staff and our users). She has also<br />

spent a considerable amount of time writing Labview code to analyze the farfield<br />

x-ray intensity patterns formed by imperfect glass capillaries, hoping that<br />

CHESS News Magazine 2009 Page 9


those patterns can be used to pinpoint the sources of error in the figure and surface perfection of the glass. Rachel Pauplis, a<br />

<strong>Cornell</strong> junior in Physics and Asian studies, worked with Irina Kriksunov, Chae Un Kim and Marian Szebenyi to study the process of<br />

pressure freezing protein crystals using a new apparatus being commissioned for MacCHESS users. Pressure freezing crystals is a<br />

fairly new technique that often results in better protein crystallography data by avoiding the need to use cryoprotectant additives,<br />

among other benefits. Rachel, a neophyte at the start, helped pinpoint necessary improvements to the apparatus and procedures<br />

which should, ultimately, make it easier to apply pressure cooling to users’ crystals on a routine basis.<br />

As you can see, this year we had a sizable group of productive<br />

<strong>Cornell</strong> students working on research projects. For the coming<br />

years, CHESS has requested support for programs that would allow<br />

us to continue to recruit <strong>Cornell</strong> students as well as those from<br />

wider regional and national pools. We look forward to continuing<br />

our summer tradition of working with enthusiastic young<br />

scientists in training.<br />

References:<br />

1. Debashis Ghosh, Jennifer Griswold, Mary Erman, and Walter<br />

Pangborn; “Structural Basis for Androgen Specificity and<br />

Oestrogen Synthesis in Human Aromatase”, Nature 457 (7226),<br />

219-223 (2009)<br />

2. The original news release can be found at: http://www.hwi.<br />

buffalo.edu/newsroom/Press_Release_09/January/1_8_09.pdf<br />

3. Published online before print February 25, 2009, doi: 10.1073/<br />

Rachel Pauplis<br />

pnas.0812071106; PNAS 106 no. 11, 4225-4230 March 17, 2009<br />

4. Read the original Rice Report here: http://www.media.rice.edu/media/NewsBot.aspMODE=VIEW&ID=12128<br />

5. Visit the Tao research group for more information: http://www.bioc.rice.edu/~ytao/index.html<br />

6. See an interesting video explanation of virus structure and function<br />

that includes quotes from eminent CHESS user Michael Rossmann: http://www.livescience.com/health/090217-virus-coat.<br />

html<br />

7. J.V. Bernier, J.-S. Park, A.L. Pilchak, M.G. Glavicic, and M.P. Miller; “Measuring Stress Distributions in Ti-6Al-4V Using <strong>Synchrotron</strong><br />

X-Ray Diffraction”, Metallurgical And Materials Transactions A 39a, 3120 (2008)<br />

8. The ACS meeting archived video news conference is here: http://www.ustream.tv/recorded/2009896<br />

9. NPR Science Friday radio interview with Jennifer Mass: http://www.npr.org/templates/story/story.phpstoryId=112105590<br />

10. Coverage of the development of the technique is here: http://news.chess.cornell.edu/articles/2007/WollWyeth.html<br />

11. Science writer Anne Ju of the <strong>Cornell</strong> Chronicle wrote the following summary (PDF): http://www.news.cornell.edu/stories/<br />

Aug09/WyethColor.html<br />

12. Lightsources.org coverage (PDF): http://www.lightsources.org/cms/pid=1003645<br />

Page 10 CHESS News Magazine 2009


Staff in Focus - John Kopsa<br />

In March of 1988 John Kopsa, CHESS<br />

Machinist, began working at the <strong>Cornell</strong><br />

Plantations mowing the grass and<br />

planting trees. A year later, a call was<br />

placed to John, asking him if he would<br />

come and hang some lead/steel panels<br />

for x-ray shielding of the experimental<br />

hutches over at Wilson Lab, to help<br />

build a new “East” research area for the<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong><br />

(CHESS). He agreed.<br />

After the panels were done, Boris<br />

Batterman (Director Emeritus), offered<br />

John a permanent position at CHESS.<br />

By the time the East area was completed<br />

it was time to rebuild the West section;<br />

that was in 1993. John was involved<br />

with the plumbing and vacuum work.<br />

John Kopsa working in the CHESS Machine Shop.<br />

Laura Houghton<br />

When the construction was finally<br />

done, John moved into the machine<br />

shop with Basil Blank and Walt Protas<br />

(Equipment Machinist). It was there that<br />

he began to do machine work. On April<br />

20th, 2000 he became a Journeyman<br />

Toolmaker. Through a correspondence<br />

course and hands on training, John<br />

completed the New York State<br />

Department of Labor requirements for a<br />

Tool and Die Maker.<br />

Except for a period of time spent<br />

assisting with the build of G-line, the<br />

machine shop is where John can still be<br />

found. His time in the machine shop<br />

has been spent building the equipment<br />

to fill the East, West, and G-line stations.<br />

He has touched almost every aspect<br />

of the experimental equipment<br />

throughout the lab.<br />

According to John, “The best part of the<br />

job has been keeping the Staff Scientists<br />

humble, especially Ken Finkelstein”.<br />

He told me that he is also looking<br />

forward to seeing what the ERL (<strong>Energy</strong><br />

Recovery Linac) will bring and how the<br />

lab will change.<br />

“I met my wife at Wilson lab. We got<br />

married in 2000 and have a fantastic<br />

young daughter on the high honor roll<br />

who loves Math & Science”, John states<br />

proudly. “We live in a big farmhouse<br />

spending a lot of time taking care of the<br />

yard and over two acres of flower gardens.<br />

Shell, my wife, and I are looking forward<br />

to riding our bikes this summer”.<br />

John enjoys many different genres<br />

of music which can almost always be<br />

heard when entering the machine shop.<br />

He is a diehard fan of the Grateful Dead<br />

as well as Johnny Cash. How’s that for a<br />

combination<br />

He takes great pride in his work, is<br />

dedicated, and is a well respected<br />

member of the CHESS team.<br />

1999<br />

2007<br />

... a few words from John’s supervisor, Dana Richter ...<br />

John has been involved in most of the<br />

large projects in the lab over the past 20<br />

years. He started out putting together our<br />

experimental hutches when he first came<br />

here in 1989. The work involved putting<br />

up large shielding panels and forming<br />

and placing concrete shielding blocks.<br />

He has also installed cooling and exhaust<br />

systems. John is our expert at fabricating and<br />

assembling optical tables and translation stages.<br />

Many people do not realize that John and<br />

the hutch assembly crew often would put<br />

little notes inside the hutch walls as they<br />

put them up and we have found some of<br />

them as we have upgraded beam lines.<br />

He now spends most of his time in the<br />

machine shop where he has become<br />

proficient with the numerically controlled<br />

milling machine. John makes many of<br />

the parts used throughout Wilson Lab.<br />

CHESS News Magazine 2009 Page 11


Facility Upgrades, Networking and Computing<br />

Chris Conolly<br />

C-line Upgrades and New Capabilities<br />

Ernie Fontes and Ken Finkelstein<br />

D-line Upgrades and New Science Capabilities<br />

Ernie Fontes and Detlef Smilgies<br />

Education & Outreach<br />

Lora Hine<br />

MacCHESS Advances: Microcrystallography & More<br />

Station F3<br />

Marian Szebenyi<br />

Ulrich Englich<br />

XPaXS: Improved Data Acquisition & Real-Time Analysis at CHESS<br />

Darren Dale<br />

CHESS <strong>High</strong>-pressure B1/B2 Update<br />

Zhongwu Wang<br />

The <strong>Cornell</strong> <strong>Energy</strong> Recover Linac:<br />

update on a future source of coherent hard x-rays<br />

Don Bilderback, Bruce Dunham, Georg Hoffstaetter, Alexander Temnykh, and Sol Gruner<br />

Facility <strong>High</strong>lights<br />

Page 12 CHESS News Magazine 2009


Facility Upgrades, Networking and Computing<br />

Chris Conolly<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Transparent upgrades<br />

“We have done so much with so little for so long<br />

we can do anything with nothing”<br />

-- author unknown<br />

While not completely true, here at CHESS,<br />

sometimes it feels that way. I like to think we<br />

could do almost anything with what we have, and<br />

sometimes we do. As I have looked over the list<br />

of projects over the past few years, a surprising<br />

number of upgrades have taken place, improving<br />

the capabilities of almost every experimental<br />

station. Some are quite obvious to the casual<br />

user of the lab, but many are not and deserve<br />

to be mentioned. Driven to innovate, the CHESS<br />

Operations staff, through a collaborative effort,<br />

make ideas into a reality. In the following article,<br />

I will describe some of the upgrades that you may<br />

or may not have heard about and who made them<br />

into a reality.<br />

Video, driving motors, racks that roll, and<br />

(hopefully) no more floods<br />

Upgrades can take on many forms and affect<br />

users in many different ways. One that improves<br />

the quality of x-rays directly is the installation of<br />

additional video beam position monitors (VBPMs)<br />

at F1 and C1 stations, which were designed and<br />

installed by operator and designer Tom Krawczyk.<br />

Viewing the F-line wiggler beam directly after it is<br />

reflected off the collimating mirror, the F1 VBPM<br />

speeds tuning by eliminating the laborious process<br />

of taking burns, which process involves opening<br />

the helium chamber, overriding the equipment<br />

protection system, and running back and forth<br />

from the cave to the control area numerous times.<br />

Diamonds have become our best friends at<br />

CHESS lately. The move towards more UHV<br />

monochromators has created a need for VBPM’s<br />

that do not rely on visible light fluorescence of<br />

helium. To the rescue are movable, thin diamond<br />

foils that can be moved into the beam as needed.<br />

They use the same video cameras and processing<br />

software, and they have proven very useful for<br />

beam alignment and even position monitoring, if<br />

the experiment can tolerate a slight loss of flux.<br />

Current systems have been installed on the D- and<br />

G-beam lines.<br />

Former CHESS Operator and current electronics<br />

designer, Eric Edwards,<br />

has made a revolutionary<br />

upgrade to how we move<br />

motors throughout the<br />

lab. In designing what he<br />

terms the MDCP -- which<br />

stands for motor drive,<br />

control, and power -- Eric has<br />

simultaneously reduced the<br />

size and cost to drive a motor<br />

while increasing the reliability<br />

and usability of the system.<br />

Now, many sets of long cables<br />

and separate components<br />

have been reduced to a<br />

single chassis, a set of motor<br />

cables, Ethernet, and power.<br />

Eric packaged a Galil motor<br />

controller and eight Gecko<br />

brand motor drives coupled<br />

to an in-house designed<br />

circuit board, all sandwiched<br />

into a 2U electronics chassis.<br />

Installations on optical tables<br />

and diffractometers have significantly<br />

decreased the time required to move<br />

experiments around the laboratory<br />

floor and get them running again.<br />

Through the efforts of Ernie Fontes,<br />

Ellen Kathan, Phil Sorenson and<br />

many others the beam line front<br />

end electronics have been slowly<br />

evolving over the past few years.<br />

Gone are the days of electronics<br />

scattered into any space<br />

available under the beam lines.<br />

Lead-shielded rolling electronics<br />

racks have been installed in the<br />

CESR tunnel near each front end.<br />

Each rack is the nerve center for its<br />

beam line -- containing ion pump<br />

controls, thermocouple monitors,<br />

motor drives, position monitor<br />

supplies, and computer networking.<br />

Tethered to the front end by a thick<br />

bundle of cables and wires, the<br />

wheels allow the rack to swing out<br />

of the way of activities and repairs<br />

taking place on CESR. Although staff<br />

must still crawl around under the<br />

The F3 XRF optical table setup can roll into<br />

place in minutes and be fully operational<br />

a short time later with its own on-board<br />

MDCP motor drives.<br />

A typical rolling rack which serves<br />

as the nerve center for each front<br />

end. This particular rack belongs<br />

to A-line and houses PE monitor<br />

electronics, ion pump controllers, a<br />

thermocouple local monitor, patch<br />

panels, a networking switch, a<br />

multiport serial to Ethernet adapter<br />

and power conditioning.<br />

CHESS News Magazine 2009 Page 13


eam lines we no longer have to troubleshoot electronics<br />

while we are there.<br />

A detector system recently installed by Ted Luddy, one<br />

of our long-time CHESS Operators, is one which users<br />

will, hopefully, never get to experience -- the flood<br />

detection system. In the past, whenever there was<br />

a breach in any of the five different water systems<br />

around the lab, the first indication of a problem<br />

was an alarm from a water sensor in the CLEO pit, the<br />

lowest point in the lab. As one can imagine, the probability<br />

that equipment damage would occur was high, leading<br />

to machine down time and experiments being delayed.<br />

The CHESS<br />

Flood Alarm.<br />

Green lights indicate<br />

that all CHESS areas are dry.<br />

Now a series of sensors have been installed in critical areas -- such as beam line front ends, optics caves, and station roofs<br />

-- to give advanced warning. The system pages the CHESS Operator directly and also sounds an alarm in the CESR control<br />

room whenever the slightest drip is detected. The system uses an off-the-shelf continuity sensor coupled with a relay box<br />

designed and built by Ted in our own electronics shop. We all sleep a little better knowing that thousands of gallons of<br />

water from one of our many water systems has less of a chance of flooding our experimental floor. Thank you for that, Ted.<br />

Another behind the scenes upgrade is a gas flow monitoring system installed by Operator Dave Jones. You may have<br />

wondered how a CHESS Operator knew you had set the gas flow to your flight tube or ion chamber a bit too high. Simply<br />

put, units have been installed for CHESS East, West, and G-line to log the pressure and flow of the helium and argon<br />

distributed gases. For example, a quick glance at the signal-monitoring web page tells the Operator whether the lab’s<br />

helium usage is normal or excessive, and then he or she can easily and efficiently make the proper adjustment. The system<br />

helps us to see trends develop and allows us to reduce usage, when necessary, saving thousands of dollars a year in costs.<br />

The compact F3 white beam mirror. Notice the<br />

external bender mechanism at the top indicating<br />

the mirror has a bounce down configuration.<br />

It’s all done with mirrors<br />

Operators Tom Krawczyk and Aaron Lyndaker traveled to the APS to use the<br />

long-trace profilometer to characterize three newly-fabricated glidcop white<br />

beam mirrors. C1 and F3 each now have an 800mm-long mirror installed in<br />

their front ends. Compact new benders for them have been designed, which<br />

locate the motor driven actuator on the exterior of the UHV mirror chamber<br />

to facilitate ease of installation and repair, if ever needed. The UHV chambers<br />

have been designed to<br />

take up the minimum<br />

of volume to be able<br />

to fit into the tight<br />

space available on the<br />

respective front ends. A<br />

spare for the aging A1<br />

collimating mirror has<br />

also been fabricated and<br />

is standing by.<br />

The F3 monochromator<br />

second optics cart as<br />

model and in reality.<br />

CHESS Engineer Alan Pauling has redesigned the F3 monochromator<br />

for better stability and increased capabilities to accept beam from<br />

the new F3 white beam mirror. The re-fit includes a new set of white<br />

beam slits located in their own upstream chamber and a longer second<br />

crystal travel stage. Ever creative, Alan used an existing port on the<br />

monochromator chamber to support the first crystal stages to allow<br />

overlap with the second crystal stage for low energy experiments. The<br />

new monochromator can use silicon or multilayer optics over an energy<br />

range of 1keV to 135keV and bandwidths up to 0.6%.<br />

Sharing space with the F3 monochromator in the large helium<br />

coffin, the F1 monochromator required some re-design to avoid an<br />

interference with the first crystal mount of the F3 monochromator. Alan<br />

designed a new compact cooled crystal bender which incorporates a<br />

more reliable bending mechanism coupled to a new copper block with<br />

Page 14 CHESS News Magazine 2009


e-routed cooling lines. During the project, the helium<br />

chamber and all its internal components were cleaned to<br />

near UHV standards and all radiation-damaged stepper<br />

motor wiring was replaced. The complete motor drive<br />

system for the 56 motors located in the F-cave was replaced<br />

with MDCPs and re-located from the cave roof to a leadshielded<br />

rack inside the F-cave. Overall, the re-fit has<br />

proven to be reliable and the additional capabilities have<br />

been put to good use.<br />

Alan Pauling and Aaron Lyndaker put the finishing<br />

touches on the C-line front end before the CESR<br />

dipole magnet is slid back into place.<br />

A phased reconstruction of C-line began during the 2009<br />

winter down period. Replacement of a section of CESR<br />

beam pipe required the removal of the CESR dipole magnet<br />

next to C-line and afforded CHESS the luxury of space for<br />

one short week to complete the installation of a new front<br />

end. A new copper-flared chamber and CESR X-ray Beam<br />

Size Monitor optics chamber were installed along with the<br />

previously mentioned white beam mirror. The C1 mono<br />

box was raised up a few inches to be able to accept the<br />

mirror reflected beam. This required the monochromator<br />

offset to change, which led to a complete overhaul of the<br />

existing monochromator, followed by motor rewiring and<br />

plumbing. Additional pink and white beam safety bricks<br />

were also installed in the coffin. Completion of phase one<br />

of this project reduces the number of beryllium windows in<br />

the beam line down to one, which will be removed in phase<br />

2 to allow the x-ray beam-size monitor (xBSM) project to<br />

gather data at low energy. The final phase will be new UHV<br />

optics, which will allow C1 to utilize low energy x-rays and<br />

continued xBSM data collection.<br />

Vacuum goes for a walk<br />

The CHESS Vacuum group, under the direction of group<br />

leader Bob Seeley, has endured many changes of venue<br />

over the past few years. Bumped from their long-time home<br />

just off the CHESS control area by the ERL project, the group<br />

has become mobile. A small fraction of the CESR vacuum<br />

lab has been turned over to CHESS for general vacuum<br />

processing needs, but larger fabrication projects have had<br />

to take place elsewhere. An additional space has been<br />

carved out of Room 128 -- a clean room built specifically for<br />

ERL development -- for clean UHV assembly work. Many of<br />

the stockpile of flanges, controls, and spare parts are now<br />

shared with the CESR group. CHESS specific components<br />

and tools now share space with the CHESS Glass Puller in<br />

RM 180. Although nomadic, productivity has increased<br />

significantly for this group with many difficult installations<br />

under their belt. The group should be commended for<br />

being flexible enough to work in this unusual situation.<br />

Multiple networks and processors<br />

On the computing front, a new private control network<br />

has been installed that is isolated from the general CHESS<br />

laboratory network. A need developed for a secure network<br />

for devices controlled over Ethernet. The implementation of<br />

Parker Compumotor motor drives and later the new MDCP<br />

drives forced the computer group to look seriously at a<br />

separate secure network. Each station has been connected<br />

to the private network to accommodate motor drive<br />

upgrades, present and future. MacCHESS has also joined<br />

the trend, installing its own local networks at A1, F1, F2 and<br />

F3. Keeping the flow of data local to each station will also<br />

have the benefit of keeping network issues localized and<br />

speeding data transfers.<br />

Out in the CESR tunnel, each beam line front end has<br />

been configured utilizing the private network to allow<br />

remote management of motor drives, ion pump controls,<br />

and, potentially, thermocouple and position monitors. By<br />

using multiport serial to Ethernet converters, devices can<br />

communicate bi-directionally with an EPICS display located<br />

at the CHESS<br />

control area, which<br />

allows operations<br />

staff to view status,<br />

trends, and even<br />

control the devices.<br />

One advantage of<br />

this system is to<br />

allow access to the<br />

hardware located<br />

in the CESR tunnel,<br />

without calling<br />

for a machine<br />

access; this reduces<br />

downtime for user<br />

experiments.<br />

Operator Ellen Kathan<br />

completed wiring one<br />

of the rolling racks in<br />

the CESR tunnel. This<br />

rack contains all of the<br />

pump controllers for<br />

the F hard bend and<br />

wiggler front ends.<br />

Consolidation of the computers that process the VBPM<br />

signal to a central location has been led by Operator Lee<br />

Shelp. Using USB converters to bring the many video signals<br />

back to the CHESS control area eliminates the need for<br />

individual computers taking up space throughout the lab,<br />

and also minimizes the potential for failure from multiple<br />

computer systems running multiple programs.<br />

CHESS News Magazine 2009 Page 15


The design group has seen many upgrades to the popular AutoCAD Inventor CAD suite over the past few years. CAD<br />

workstations have all been upgraded to dual-processor machines running 64-bit Windows to utilize all of the functionality<br />

of the program. This upgrade also benefits ANSYS users, who are now able to churn out simulations at an ever-increasing<br />

rate.<br />

Smaller is always better in the CHESS world. This is certainly the case when it comes to diagnosing and testing motor<br />

problems quickly. Operations staff now have a few netbooks available to allow one person to single-handedly move<br />

motors locally rather than requiring a second person to run them remotely. Using an ssh session over the wireless<br />

network, operators can quickly locate and repair motor problems in the caves and CESR tunnel. These basic machines are<br />

inexpensive, lightweight, and run a variety of operating systems. They have also found a use running RGAs and viewing<br />

signals using a specially developed fast-monitoring windows program.<br />

New Blood<br />

The changes around the lab aren’t simply with systems upgrades and<br />

modifications. We have added some new faces to our staff as well.<br />

Within the past year, we have hired two new CHESS Operators. Rich Jayne<br />

jumped right in with both feet to design, fabricate, and install a MAR345<br />

detector cradle in the B2 station. The overhead rail system gives ample<br />

clearance for experiments, while allowing effortless motion in the x, y,<br />

Operator Rich Jayne inspects<br />

the MAR 345 detector in the<br />

B2 station.<br />

and z directions as well as in rotation.<br />

Solenoid actuated brakes lock the<br />

detector firmly in place. A less involved<br />

mount for our other MAR345 detector<br />

was requested for the A2 station on short<br />

notice and Rich delivered -- cobbling<br />

together a motorized table from spare<br />

parts found around the lab. Alignment time of the detector for the Miller<br />

group has decreased by an order of magnitude and allowed the group to<br />

collect data sooner in their allotted beam time. A positive outcome for all.<br />

Our other new Operator, Aaron Lyndaker, has brought excellent design skills<br />

to the group. Beam line front-end design at CHESS requires that we be able<br />

to think and work in very tight spaces. In most cases, removing components<br />

from a beam line makes this easier, but when removing beryllium windows this<br />

isn’t the case. Once the beryllium windows are removed, a differential-pumping<br />

section must be installed to protect CESR from the higher pressure at the experiment.<br />

Differential pumping requires additional valves, pumps, and apertures to work. Aaron was able to incorporate all of these<br />

into a clean design for D-line. So far, our first foray into windowless beam lines has been a success, allowing the CESR xBSM<br />

experiment to happen in a very short time frame. As part of this project, the old D-line monochromator was scrapped<br />

and a new UHV set of optics, designed by Tom Krawczyk, was installed. The detector used by the xBSM experiment is not<br />

UHV compatible, so a single, very thin 5 micron diamond window was purchased to separate the detector chamber from<br />

accelerator vacuum. Using a complex pressure-balancing system, designed by Engineer Jim Savino, the pump-down and<br />

venting of the xBSM detector chamber has been automated using a PLC. A single button press allows the user to access<br />

the detector at any time while protecting the UHV vacuum system connected directly to the accelerator vacuum.<br />

There is also a new face in the computer group, Shijie Yang, who will help with hardware and software maintenance and<br />

system administration. He will also help train and supervise staff, students, and visitors in the many software packages<br />

used around the lab.<br />

These are but a few of the upgrades CHESS has seen over the past few years. Many thanks for the countless hours spent by<br />

the tireless CHESS staff that has resulted in less downtime and increased capabilities. Who knows what future upgrades are<br />

in store for CHESS but one thing is certain, the CHESS staff will be there with creativity and a drive to make them a reality.<br />

Page 16 CHESS News Magazine 2009


C-line Upgrades and New Capabilities<br />

Ernie Fontes and Ken Finkelstein<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

This article will highlight developments at C-line, including recent upgrades, new science capabilities and long-term plans.<br />

Recall that C-line is a bend-magnet-source beamline that features very flexible x-ray optics and has one of the smallest x-ray<br />

beam source sizes of any CHESS beamlines. Primary use of the beamline has been for energy dependent resonant (anomalous)<br />

scattering including 4-circle diffraction, anomalous SAXS, EXAFS and XANES. In addition the station is used for x-ray topographic<br />

studies, multiple beam diffraction, high pressure x-ray spectroscopy, and specialized x-ray optics development. See article on<br />

page 58 by Richard Jones – Topography of Diamonds at CHESS helps Nuclear Physics Program at JLAB.<br />

Several years ago, while writing a proposal for CHESS<br />

funding starting in 2008, the scientific staff identified<br />

research capabilities that were missing from CHESS and<br />

new scientific communities they would benefit. At the top<br />

of the list was the need to extend the x-ray beam energy<br />

range down to 2 keV to accommodate low energy x-ray<br />

applications. This will open new opportunities in biology,<br />

condensed matter and environmental sciences and give<br />

the CHESS staff invaluable experience in a new spectral<br />

region. A unifying theme of the new scientific capabilities<br />

was the need to perform resonant elastic x-ray scattering,<br />

anomalous SAXS, crystallographic phasing, solution x-ray<br />

absorption (XANES), and to study quantum correlations by<br />

gaining access to low-energy absorption edges. If C-line<br />

could reach down to 2 keV, figure 1 shows the elemental K,<br />

L, and M edges which would then become accessible.<br />

In the past, CHESS has not supported x-ray applications<br />

below 6 keV because our beamline designs traditionally<br />

Fig. 1: Periodic table showing low-energy absorption edges that will become<br />

newly accessible at the C1 station after technical upgrades. Blue, red, and green<br />

indicate K,L,M edges, respectively.<br />

included two beryllium vacuum windows, typically 0.01<br />

inch ultrapure foils. The engineering purpose of these windows was to separate the downstream x-ray optics box atmosphere,<br />

typically helium, from the sensitive upstream ultra-high-vacuum environment needed for the particle accelerator. While<br />

providing vacuum protection, these windows also dramatically attenuate photons with energies below 6 keV, making it<br />

impractical to collect data at light element absorption edges, for example. To overcome this limitation and improve the<br />

capabilities of C line, the engineering staff outlined a plan summarized in figure 2. It involves upgrading the beamline front-end,<br />

beam transport, x-ray optics, and experimental hutch components to be compatible with (clean) vacuum equipment and to<br />

include differential pumping segments, and fast-acting sensors and valves to protect against component failures. With upgraded<br />

pumping and protective valves in place, we can remove all windows during normal x-ray operations.<br />

Fig. 2: Schematic plan view of the major<br />

upgrades proposed for the C-line beamline<br />

and C1 station. At left, a new white beam<br />

vertical-focusing 800 millimeter glidcop mirror<br />

was installed and used. Placeholders for a<br />

low-energy mirror and new high-vacuum x-ray<br />

optics box are show. Long term we hope to<br />

extend the length of the hutch: shielding wall<br />

positions indicate “old” (dashed) and “new”<br />

positions. The new positions would extend the<br />

hutch approximately two meters so that the<br />

diffractometer could be located at a 1:1 focusing<br />

arrangement. Not shown off the left side is a<br />

small vacuum box to hold the CESR-TA pinhole<br />

camera optics.<br />

CHESS News Magazine 2009 Page 17


In addition to new capabilities for low-energy x-ray<br />

experimentation, removing windows from the beamline also<br />

makes it possible for a group of accelerator physicists to study<br />

the operation of the storage ring at lower beam energy, e.g.<br />

2.0 GeV. The CESR-TA group, as they are called, is using CESR<br />

as a test accelerator and model of the positron damping ring<br />

proposed for the International Linear Collider project 1 . In their<br />

studies, the C beamline and station will be used to create a pinhole<br />

camera to view the electron bunches stored in CESR. The<br />

pin-hole camera is formed by putting a small slit or pin-hole in<br />

the tunnel, upstream of the C-line shielding wall, and placing a<br />

diode array detector in the downstream-most station position.<br />

Since the work is done with a low-energy particle beam (2<br />

GeV) the windowless arrangement of the beamline is needed<br />

to maximize the photon flux for their measurements. These<br />

measurements will be done outside the x-ray run schedule, and<br />

so will not affect x-ray users at C1. A similar beam-size monitor<br />

was commissioned last year on D-line to study positron beams;<br />

see the discussion of the D-line upgrades in a separate article.<br />

The C-line upgrades included an extensive rearrangement and<br />

replacement of upstream vacuum components over the course<br />

of two CESR down periods. In the upstream-most position,<br />

a new copper-flared chamber was installed to contain and<br />

aperture the x-ray beam when it first exits the storage ring.<br />

Just downstream, a new small vacuum chamber was added to<br />

allow the CESR-TA group to install apertures and pinholes for<br />

source imaging experiments. During x-ray running periods,<br />

this pinhole optics is completely withdrawn and does not<br />

interfere with the x-ray user beam. Downstream of the main<br />

line beamstops, pieces were installed to accommodate a new<br />

white beam vertical focusing mirror. Fabricated by CHESS<br />

with help from LBNL, this new 800 millimeter-long glidcop<br />

mirror is internally-cooled and located for 1-to-1 focusing of<br />

the CESR source at the center of the C1 diffractometer in the<br />

experimental hutch. This mirror will improve C line in four<br />

ways: 1) focusing will increase flux density (photons/second/<br />

area) at the sample by about 3 times, without increasing<br />

angular divergence or reducing energy resolution. 2) The mirror<br />

can be used for (vertical) angle collimation to produce a more<br />

parallel x-ray beam. This reduces vertical beam size at the<br />

sample and increases spectral flux density (photons/second/<br />

unit bandwidth). 3) The mirror also reduces total power on the<br />

monochromotor first crystal, typically by a factor of 2. 4) It can<br />

be adjusted to remove high energy harmonics of the primary<br />

beam.<br />

To satisfy all the goals of the project we need to completely<br />

replace the x-ray optics box, but for the running periods in 2009<br />

we only had time to modify existing parts to accommodate<br />

variable height of mirror reflected beam. These modifications<br />

included redesigning the monochromator to permit height<br />

adjustment of the first crystal, and installation of a novel beam<br />

viewer system downstream of the crystals. The viewer uses<br />

a 40 micron thick, 100 mm diameter silicon wafer to convert<br />

x-rays to visible light. Light from the crystal is reflected by a<br />

mirror to a video camera. The thin crystal will be a very useful<br />

diagnostic and alignment tool because it can intercept filtered<br />

white beam, mirror reflected “pink beam” and/or image Laue<br />

diffraction from the first crystal. These capabilities allow us<br />

to use the x-ray beam for precise positioning of the mirror<br />

and mono first crystal. We have also added “phi-adjustment”<br />

(rotation about the surface normal) to the first crystal. Phiadjustment<br />

is critical when using asymmetrically cut crystals<br />

(diffraction planes NOT parallel to the surface), for example<br />

to expand the beam for topography studies (as was needed<br />

during the topography work by Richard Jones). The support<br />

structure for the second mono crystal has also been rebuilt for<br />

greater stability and for operation at fixed height above the<br />

white beam. To accommodate these improvements we have<br />

also rebuilt the white beam slit system for the wider range of<br />

vertical beam position, raised up the optics box to match the<br />

new range of beam positions, built new apertures to ensure<br />

safe handling of white and pink beams, and added a capability<br />

for white beam use in C-hutch by the accelerator physics group.<br />

Work is also ongoing to design new instruments for the<br />

experimental hutch to meet goals to expand science<br />

capabilities at C line. One project involves design of an x-ray<br />

emission spectrometer for collecting x-ray fluorescence<br />

with high energy resolution over a large solid angle 2 . This<br />

capability is being built in collaboration with Serena DeBeer<br />

George 3 , a new <strong>Cornell</strong> chemistry professor who is interested<br />

in development and application of x-ray based spectroscopies<br />

to probe electronic structure in biological and chemical<br />

catalysis. In particular, she studies transition metal active<br />

sites that are often centers of catalytic activity. Active sites<br />

provide inspiration for biomimetic model chemistry and for<br />

understanding homogeneous catalysis. Once the emission<br />

spectrometer is perfected it will become available to all CHESS<br />

users. The present design will collect about 10 times more solid<br />

angle than is possible with our large aperture single element<br />

solid state detectors (e.g Vortex), and it will provide energy<br />

resolution matched to the width of emission<br />

lines (see figure 3). This will allow collection of:<br />

weak emission, closely spaced lines, and lines<br />

that are very close in energy to the incident<br />

beam. We plan to couple this spectrometer<br />

to a fast readout, low noise position sensitive<br />

detector to permit simultaneous data collection,<br />

over a range of energy sufficient for line shape<br />

analysis.<br />

Continued on pg. 21<br />

Page 18 CHESS News Magazine 2009<br />

Fig. 3: Schematic design for the high-efficiency triple-crystal x-ray fluorescence<br />

detector. The incident x-ray beam (from left) excites fluorescence in the specimen, and<br />

that light is collected by three spherically-bent silicon crystals and focused to a highcount-rate<br />

photon counting detector. Shielding and helium flight paths are not shown.


D-line Upgrade and New Science Capabilities<br />

Ernie Fontes and Detlef Smilgies<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

CHESS’s popular D1 station, utilizing synchrotron radiation<br />

from a hard-bend dipole magnet and a multilayer<br />

monochromator, has both a versatile optical system and a<br />

5-meter long hutch (Figure 1). Sharing the smallest x-ray<br />

source size at CHESS with C-line, D1 has been extensively<br />

used for small and wide angle scattering as well as<br />

microbeam experiments with glass capillaries to create x-ray<br />

beams on the scale of 50 nanometers to 50 microns. Recent<br />

applications include grazing-incidence small- and wideangle<br />

x-ray scattering on thin films of soft materials such<br />

as block copolymers 1 , conjugated polymers 2 , nanocrystal<br />

assemblies 3 and nanoporous films 4 as well as high-energy<br />

small-angle scattering in conjunction with diamond anvil<br />

cells 5 , and microbeam SAXS and WAXS studies 6 . D1 has also<br />

supported a development program for new experimental<br />

techniques such as an evaluation of the potential of<br />

Laue diffraction for structure determination of protein<br />

microcrystals 7 or fluorescence imaging of large-scale<br />

paintings 8 and Roman tombstones 9 .<br />

At about the same time the accelerator physics<br />

group was proposing to use the CESR storage ring<br />

as a “test accelerator” (TA) for part of the R&D aimed<br />

at the international linear collider (ILC). Using the<br />

flexible lattice of CESR the TA group wanted to<br />

create a prototype ultralow emittance positron<br />

Fig. 1: Conceptual plan to improve the D1 experimental program by lengthening<br />

the user space by 2 meters (old wall position shown). The new high-vacuum x-ray<br />

optics box is already installed and commissioned, and D1 has delivered a total of<br />

12 weeks of user operation within the old experimental hutch (dashed).<br />

A few years ago, during an exercise to identify strategic<br />

long-term plans to improve CHESS stations and beamlines,<br />

it became clear that the size and scope of the soft-matter<br />

science program being supported at D1 was “bursting at<br />

the seams.” For example, new demand has been growing<br />

steadily to study thin-films either in static specimens or<br />

in-situ during solvent vapor conditioning. Technically,<br />

the beamline front-end and optics had not been<br />

upgraded in almost two decades. In particular, the helium<br />

monochromator box that housed the synthetic multilayer<br />

mirrors was in poor shape and the unclean environment<br />

led to carbon-like deposits on the multilayer surfaces that<br />

needed to be cleaned off periodically.<br />

This exercise resulted in replacing the old helium optics<br />

box with a high-vacuum enclosure and a conceptual design<br />

to lengthen the experimental hutch (see figure 1). The<br />

extra space became available with the demise of the high-<br />

energy physics program at the decommissioned<br />

CLEO detector. The longer hutch would allow<br />

access to SAXS studies with higher resolution<br />

and/or at higher beam energies. A new optics<br />

box was designed to provide rapid changeover<br />

between multiple multilayer optics with different<br />

wavelength ranges, thus cutting down on set-up<br />

and realignment time.<br />

and electron damping ring of the sort that might become part of<br />

the injector system for the ILC. In this study they needed to setup<br />

a pinhole camera to image the electron and positron bunches,<br />

using the CHESS x-ray beamlines C1 and D1 close to the former<br />

interaction region. Using a fast detector, the group hoped to<br />

be able to measure the size and shape of individual bunches of<br />

charged particles, with photons emanating from a single pass<br />

of the bunches through the bend magnet source. Because this<br />

work would be done with a machine energy of 2.0 GeV, where the<br />

x-ray flux is peaked at 1 keV, the accelerator physicists requested<br />

that the targeted beamlines have all beryllium vacuum windows<br />

removed, in order to avoid absorption of low-energy x-rays<br />

coming down the line.<br />

Marrying the needs for better x-ray science capabilities with the<br />

accelerator group requests, the CHESS design, technical and<br />

vacuum groups designed an upgrade plan to accomplish these<br />

goals. Pictured in figure 2, the upstream D beamline was almost<br />

completely replaced with ultra-high-vacuum compatible flight<br />

tubes, apertures, a differential pumping stage to separate storage<br />

CESR hard bend dipole magnet<br />

quadrupole<br />

CESR-TA<br />

optics box<br />

D1 optics box<br />

Fig. 2: The new windowless D1 front-end. The x-ray beam enters from the right, traveling<br />

downstream to the left through a series of beamstops, a beam viewer, and differential<br />

pumping chambers. The D1 optics box is described in more detail in the text and in Figure 3.<br />

CHESS News Magazine 2009 Page 19


ing ultra-high vacuum from the high vacuum typically achieved<br />

in optics boxes, as well as removable thin diamond foil x-ray<br />

beam viewers to assist line-up.<br />

For the CESR-TA project a small vacuum optics box as close to<br />

the storage ring as possible was added. This small box is used<br />

by the accelerator group to hold a small pinhole, Fresnel zone<br />

plate, or coded aperture that will create an image of the positron<br />

source. During these special “TA” runs, all D1 hutch equipment<br />

(two full CHESS optical tables) is removed and the CESR-TA<br />

group rolls into the x-ray station a large vacuum detector box<br />

that has an ultra-thin diamond window with a pressure balance<br />

system designed by CHESS. This box houses a high-speed pindiode<br />

linear array detector that measures the projected size of<br />

the x-ray source.<br />

The old D1 x-ray optics box was replaced by a high-vacuum<br />

chamber with all vacuum-compatible motors, a large<br />

turbomolecular pump, gate valves before and after the optics<br />

for isolation of vacuum cells, a set of adjustable white beam<br />

slits, and a second downstream thin diamond foil x-ray beam<br />

viewer (figures 2 and 3). When the x-ray beam passes through<br />

the diamond foil beam viewer it creates a visible light image that<br />

is recorded by a video camera. Coupled with custom software<br />

designed by CHESS scientist Peter Revesz, the in-situ video<br />

images of the x-ray beam are extremely helpful for diagnostics<br />

on beam shape, position, and stability.<br />

The new D1 multicrystal monochromator contains three pairs<br />

of multilayers of varying properties. The first set is the tried and<br />

true 30Å Mo:B 4<br />

C multilayer made by the APS optics group that<br />

has been D1’s workhorse monochromator over the past four<br />

years 10 . Although somewhat damaged by prolonged operation<br />

in the old helium box, there are still good parts left. In the last<br />

running period of Spring 2009, the x-ray beam remained at a<br />

beam<br />

limiting<br />

slit set<br />

optional<br />

mirror<br />

long translation<br />

x table<br />

multi<br />

layer 2<br />

multi<br />

layer 1<br />

single spot, and it showed no signs of degradation over an<br />

initial four week user-mode operations period. Typically this<br />

multilayer covers the 8 keV to 15 keV energy range, but can<br />

get as low as 6 keV at increased offset. The second multilayer<br />

set is again a matched set, 21Å W:B 4<br />

C by Osmic. This<br />

multilayer was chosen to cover the hard x-ray range from 15<br />

keV to 30 keV. Originally conceived for fluorescence imaging,<br />

it is now also being tested for high-pressure SAXS and GIWAXS<br />

for an extended scattering range.<br />

The third multilayer set is an experimental one for exploring<br />

new optics configurations. The d-spacings of the Mo:B 4<br />

C<br />

multilayers were purposely mismatched with 21 Å for the<br />

upstream multilayer and 25 Å for the downstream multilayer,<br />

resulting in a beam that is somewhat bent down from the<br />

horizontal. A third optical element, a rhodium-coated mirror,<br />

is used to bend the beam back up to the horizontal plane,<br />

working close to the Rh critical angle. Since multilayer<br />

scattering angles and the critical angle of the mirror both<br />

scale inversely proportional with energy, this configuration<br />

can be maintained throughout the energy range of 10-20 keV.<br />

The Pd mirror generates an appropriate energy cut-off for<br />

higher harmonics, so that the third and higher orders cannot<br />

pass. This will result in a D1 beam with much less higherenergy<br />

contamination.<br />

As an additional option, the mirror can be slightly detuned<br />

towards lower incident angles, resulting in a slightly bent<br />

down beam that could enable grazing incidence experiments<br />

from liquid surfaces. Again, as multilayer Bragg angles and<br />

relevant critical angles have the same energy dependence,<br />

this mode can be maintained throughout the energy range. A<br />

future upgrade could add a bender to this mirror to increase<br />

flux on the samples at the cost of a small extra angular<br />

divergence, while maintaining a well-defined incident angle<br />

on samples. The advanced mirror options still remain to be<br />

commissioned at the time of writing.<br />

The multilayer set-up rests on a solid internal x-translation<br />

table so that different multilayers can be brought into the<br />

beam via remote computer control. This avoids the problem<br />

of having to re-align the optical tables in the hutch after<br />

changing multilayers. All monochromator motions use<br />

vacuum-compatible motors and bearings. Vacuum of better<br />

than 10 nanoTorr was maintained under beam and motor<br />

moving after careful conditioning. The whole monochromator<br />

assembly can be leveled by an external 3-point actuator set.<br />

Additional optical elements are a set of white beam slits<br />

limiting the incident beam to the beam that is needed.<br />

These internally water-cooled slits are controlled by external<br />

actuators via bellows, and maintain a clean vacuum even<br />

under heat load. Finally, a set of fixed slits behind the<br />

multilayer optics will be used to clean up the beam exiting<br />

the monochromator; in particular, to remove small amounts<br />

of diffuse scattering from the multilayers. This slit set, to be<br />

installed before the upcoming run, is also controlled through<br />

an external actuator via bellows.<br />

3-point mount for leveling<br />

Page 20 CHESS News Magazine 2009<br />

Fig. 3: Inside detail of the new D1 vacuum multicrystal<br />

monochromator. Details given in text.


One aspect of the upgrade that has not yet been tackled is to lengthen the experimental hutch. This will require a significant<br />

restructuring of the thick and heavy shielding walls, as well as moving some quantity of wiring. Short of moving those walls,<br />

though, the downstream-most wall of the hutch had a 12-inch hole bored through it. This was done as an optional, alternate<br />

means to locate the “TA” beam size detector as far as possible from the storage ring source. This option has not yet been exploited.<br />

Another option under discussion is adding a full second D-line hutch at the far end of the CLEO enclosure, that would serve for<br />

USAXS, microbeam scattering, or coherent scattering applications as well as for CESR-TA at higher resolution.<br />

We thank the members of the CHESS Mechanical Design Group headed by Jim Savino, the Operations Group headed by Chris<br />

Conolly, and the Vacuum Group headed by Bob Seeley for all their great work. Particular thanks to Aaron Lyndaker who designed<br />

the bulk of the vacuum system including the new monochromator box, and Tom Krawczyk who designed, assembled, and<br />

commissioned the complex insides of the multi-crystal monochromator.<br />

References:<br />

1. Christine M. Papadakis, Zhenyu Di, Dorthe Posselt, and Detlef-M. Smilgies; “Structural Instabilities in Lamellar Diblock Copolymer<br />

Thin Films During Solvent Vapor Uptake”, Langmuir 24, 13815-13818 (2008)<br />

2. Mingqian He, Jianfeng Li, Michael L. Sorensen, Feixia Zhang, Robert R. Hancock, Hon Hang Fong, Vladimir A. Pozdin,<br />

Detlef-M. Smilgies, and George G. Malliaras; “Alkylsubstituted Thienothiophene Semiconducting Materials: Structure – Property<br />

Relationships”, J.Am. Chem. Soc. online (2009)<br />

3. Danielle K. Smith, Brian Goodfellow, Detlef M. Smilgies, and Brian A. Korgel; “Self-Assembled Simple Hexagonal AB2 Binary<br />

Nanocrystal Superlattices: SEM, GISAXS and Substitutional Defects”, J. Am. Chem. Soc. 131, 3281–3290 (2009)<br />

4. Sivakumar Nagarajan, Mingqi Li, Rajaram A. Pai, Joan K. Bosworth, Peter Busch, Detlef-M. Smilgies, Christopher K. Ober,<br />

Thomas P. Russell, and James J. Watkins; “An Efficient Route to Mesoporous Silica Films with Perpendicular Nanochannels”, Adv.<br />

Mater. 20, 246–251 (2008)<br />

5. Zhongwu Wang, Ken Finkelstein, and Detlef Smilgies; “Exploring New Physics of Nanoparticle Supercrystals by <strong>High</strong> Pressure<br />

Small Angle X-ray Diffraction”, CHESS News Magazine, pg 56 (2009)<br />

6. Jessica S. Lamb, Sterling Cornaby, Kurt Andresen, Lisa Kwok, Hye Yoon Park, Xiangyun Qiu, Detlef M. Smilges, Donald H.<br />

Bilderback, and Lois Pollack; “Focusing Capillary Optics for use in SAXS Solution Scattering”, J. Appl. Cryst. 40, 193–195 (2007)<br />

7. S. Cornaby, D.M.E. Szebenyi , D.-M. Smilgies , D.J. Schuller , R. Gillilan , Q. Hao, and D.H. Bilderback; “Feasibility of One-Shot-per-<br />

Crystal Structure Determination using Laue Diffraction”, (to be published)<br />

8. A.R. Woll, J. Mass, C. Bisulca, M. Cushman, C. Griggs, T. Wazny, and N. Ocon; “The Unique History of the Armorer`s Shop: an<br />

application of confocal x-ray fluorescence microscopy”, Studies in Conservation 53, 93-109 (2008)<br />

9. J. Powers, N. Dimitrova, R. Huang, D.-M. Smilgies, D.H. Bilderback, K. Clinton, and R.E. Thorne; “Recovering Ancient Inscriptions<br />

by X-ray Fluorescence Imaging”, Zeitschrift für Papyrologie und Epigraphik (Bonn, Germany) 152, 221-227 (2005)<br />

10. Alexander Kazimirov, Detlef-M. Smilgies, Qun Shen, Xianghui Xiao, Quan Hao, Ernest Fontes, Don H. Bilderback, Sol M. Gruner,<br />

(continued from pg 18)<br />

In conclusion, during this upcoming Fall 2009 x-ray running period the technical and scientific staff at CHESS will<br />

commission new beamline, optics and station equipment that promise to significantly update and enhance the scientific<br />

capabilities of the flexible C-line experimental station. Our ambitious plans to produce low-energy x-ray optics and expand<br />

the size of the experiment hutch are still being developed. Look for more innovations during the coming year and contact<br />

the station scientist, Ken Finkelstein, with any questions or suggestions.<br />

References:<br />

1. See for CESR-TA: http://www.news.cornell.edu/stories/Aug09/cesrTA.html<br />

2. U. Bergmann et.al., Chem.Phys.Lett. 302, 119 (1999)<br />

3. http://www.chem.cornell.edu/faculty/index.aspfac=70<br />

CHESS News Magazine 2009 Page 21


Education and Outreach<br />

Lora K. Hine<br />

<strong>Cornell</strong> Laboratory for Accelerator-based Sciences and Education,<br />

<strong>Cornell</strong> <strong>University</strong><br />

I have the best job in the entire laboratory. Hands down. I have never<br />

made that proclamation out loud or even dared to whisper it to anyone<br />

sitting next to me; I am afraid people will finally catch on and realize<br />

that the work I do is much more enjoyable than what they are doing.<br />

Here’s why: I do not have to follow any taxing laboratory protocol; my<br />

instrumentation does not require fine-tuning; my results seldom (okay,<br />

never) end up in any peer-reviewed publications. My office resembles a<br />

toy store, where visitors spend more time scanning the shelves crammed<br />

full of useful materials (i.e. equipment) than they do maintaining eyecontact.<br />

My trip record forms display regional travel destinations<br />

that end in the words “school district” versus distant, more expensive,<br />

locations ending with “national laboratory”. My calendar is peppered<br />

with a variety of exciting events; camps, tours, teacher workshops, lab<br />

activities/demonstrations, and even the much anticipated student field<br />

trips. My work day includes interactions with a range of people as young<br />

as third grade school children to as senior as emeritus faculty members.<br />

My job, directing outreach activities at CLASSE and helping fulfill the<br />

broader impact criterion established by the NSF, is a truly remarkable<br />

appointment.<br />

One of the most rewarding aspects of this position is the time I get to<br />

spend with young children. They are so anxious to learn something, just<br />

about anything, about science. Before reaching puberty, children are<br />

blessed with a sense of awe and wonder; their minds are wide open and<br />

waiting to be filled. To a ten year old, nothing is cooler than looking at<br />

clusters of dead skin cells underneath a microscope! Discovering that<br />

Middle school science teachers Mary Anne Ritinski (IS 234<br />

Brooklyn) and Crisan Berina (IS 126 Queens) work together<br />

during the Summer Institute to build their own light bulb to<br />

observe the effects of resistance in a wire.<br />

you can blow a bubble using a toilet-paper tube is<br />

sheer ecstasy. Donning safety goggles to witness<br />

the shattering of a flower after being submerged in<br />

liquid nitrogen is, to an elementary student, as good<br />

as life will ever get. After visiting with Wilson Lab as<br />

part of a class field trip, I have no doubt in my mind<br />

that these children run home and enthusiastically<br />

tell their family members about all of the fascinating<br />

science stuff they saw (and did!) that day. And, more<br />

importantly, they tell their parents and themselves<br />

that they want to learn more.<br />

During the last year, the laboratory has provided<br />

after-school enrichment programming for<br />

nearly 100 elementary school students. That’s<br />

a lot of kids receiving one-on-one, hands-on<br />

science programming outside of school hours!<br />

This programming is provided free of charge to<br />

any school within a sixty-mile radius of Ithaca.<br />

Many schools have children from disadvantaged<br />

backgrounds; their families live in remote, rural<br />

areas and they haven’t even heard of <strong>Cornell</strong><br />

<strong>University</strong>. Most programs run one hour a week<br />

after school for six weeks, which is just enough time<br />

for me to get to know the kids, figure out what they<br />

want to learn more about, and (hopefully!) inspire<br />

them to seek out science and be comfortable with<br />

it - not scared of it. The most popular program we<br />

deliver to schools on a regular basis is “Atoms for<br />

Kids!”, but “The Science of Bubbles” comes in a close<br />

second. During the Atoms program, students get<br />

to operate light microscopes, investigate crystal<br />

formation, study light spectra, and witness what<br />

happens to atoms inside of a bell jar. All of which is<br />

extremely cool stuff when you are a 3-5th grader….<br />

and still cool stuff when you are an adult.<br />

If only we could bottle the excitement of a fifth<br />

grade class and pass it along to class of high school<br />

students. By the time a typical high school physics<br />

class makes a trip to Wilson Lab, they are less<br />

enthralled by the wonders of science and are more<br />

concerned with grades, college, dating and a host<br />

of other distractions. This is not to say that these<br />

students are disinterested in science, but they are<br />

less enamored by its mysteries and more concerned<br />

with completing college applications! Some<br />

students have a genuine interest in conducting<br />

research, and it is the Lab’s good fortune to be able<br />

to provide some research experiences for high<br />

school students. During the past year, a number<br />

Page 22 CHESS News Magazine 2009


of high school students have worked alongside our laboratory’s<br />

scientists, involved in projects of value to their mentors. One<br />

CHESS researcher, Richard Gillilan, worked with Ithaca <strong>High</strong><br />

School student Sohyun (Sarah) Kim during the summer of 2007<br />

to help her write a Java program to analyze 100,000 digital<br />

images of crystal growth. This past summer, she gave a poster<br />

presentation on her work at the CHESS user meeting. Currently<br />

Sarah is volunteering in MacCHESS’s lab growing protein crystals<br />

for use in beam-line testing.<br />

In addition to structured internships, the laboratory hosts a large<br />

number of visiting high school students as part of class field trips<br />

to Wilson. Many science teachers, from as far away as Buffalo NY,<br />

visit the lab as part of their one (and usually only) field trip during<br />

the school year. On average, 250 high school students each year<br />

amble along the passageways of the laboratory, gawking at the<br />

massive collection of electronic gadgetry, dodging the occasional<br />

drip of tunnel-juice, and timidly asking questions about the<br />

scope of the research being done here. A tour of the facility just<br />

can’t but help some high school students feel a sense of awe<br />

and wonder about the opportunities that lie in store for them as<br />

budding young scientists.<br />

Having spent six years as a middle school science teacher, I have<br />

grown comfortable interacting with children at this age level<br />

and even more at ease with the teachers who choose to make a<br />

profession out of educating this age-group. Consequently, the<br />

bulk of the professional development opportunities available<br />

to teachers through the outreach program at the Laboratory<br />

are geared towards 6-8th grade science teachers. Contrary to<br />

popular belief, middle school students are wonderful to have in<br />

the science classroom. They are capable, curious and internally<br />

motivated to “do science”. They adore hands-on activities and<br />

investigations; they benefit tremendously from interactive<br />

experiences, and are not yet afraid to propose an explanation<br />

for observed phenomena. Much anecdotal evidence suggests<br />

that middle school is the time when students decide, somewhat<br />

subconsciously, that they enjoy or despise science. Knowing this,<br />

it is important that adults, especially teachers, fuel a student’s<br />

passion for science before they leave middle school. If not<br />

properly fueled, that passion may not ever be reignited.<br />

One can see how important it is that middle school science<br />

teachers have the capacity and training to confidently teach their<br />

students. The outreach program at Wilson lab has taken an active<br />

role in helping to provide science teachers with the professional<br />

Lincoln Street Elementary School<br />

(Waverly, NY)<br />

Students have a good time learning about<br />

surface tension and experimental<br />

design by blowing bubbles with various<br />

house-hold objects.<br />

African Road Elementary School<br />

(Vestal, NY)<br />

Students look through the spectroscopes they built<br />

to observe the different colors of light released by<br />

heated gas bulbs.<br />

development training they need by hosting the <strong>Cornell</strong> Physical Science Summer Institute for Middle School Science Teachers<br />

during the summer of 2008. The theme for the week-long institute was “Making Connections: What Science Research and Your<br />

Middle School Curriculum Have in Common”. The Institute was partially funded through grant money received from the New York<br />

State Education Department; the other funding came from money devoted to Outreach at CLASSE. The Institute, held at <strong>Cornell</strong><br />

for a week in July, provided eleven middle school science teachers from New York City with the opportunity to gain content<br />

knowledge aligned with NY State learning standards. It also provided teachers with the chance to interact with scientists who<br />

are conducting cutting-edge research and allowed participants to learn about state-of-the-art technology in the unique setting<br />

of a world-class research university. All participants earned one unit of graduate credit in Physics for the five days they spent on<br />

campus. A quote from one participant asked to reflect upon the week states; “My experience at the Physics Summer Institute was<br />

very exciting and informative. The hands-on explanations and the lectures from various experts in the field of physics can assist<br />

me in enhancing my knowledge in science as well as in delivering more research-based activities.”<br />

In addition to working with K-12 teachers and students, I am given the opportunity to interact with undergraduate students<br />

who spend time conducting research at our laboratory. These students, from diverse backgrounds and experience levels,<br />

are competent, motivated and frequently amaze me with their intellectual maturity. The science and engineering students<br />

participating in the Research Experience for Undergraduate (REU) program are selected from a large pool of applicants from<br />

CHESS News Magazine 2009 Page 23


throughout the country and come to Wilson Lab to work with faculty, research associates, and graduate students on a broad<br />

selection of research projects. These projects contain important elements of the overall research program at CLASSE; topics<br />

range from accelerator physics to microwave superconductivity to applications of synchrotron radiation in scientific research.<br />

During their final presentations delivered during the last week of the program, students showcase the complex research<br />

they have done and reveal the advanced level of thought and understanding required to work on projects alongside their<br />

<strong>Cornell</strong> mentors. Although these students work hard, they also play hard and take advantage of many social opportunities<br />

made available to them as part of the summer program. These college students live at the same dorm and hang out with REU<br />

students from other nationally sponsored programs at <strong>Cornell</strong>. REU’ers can be found playing soccer, bowling, dining, riding<br />

bikes, or sightseeing in downtown Ithaca with their colleagues throughout the summer months.<br />

During the summer of 2008, seven of the ten REU students worked for ten weeks on substantive ERL-related research projects,<br />

the efforts of which will be referenced and utilized by future staff and students. For instance, REU student Justin Hugon worked<br />

with research scientist Don Bilderback and technician Tom Szebenyi to reduce the vibration that occurs during current x-ray<br />

capillary optics fabrication. Improvements in capillary design will be necessary to take full advantage of the synchrotron<br />

radiation generated by the ERL. Upon completion of the program Justin, currently a physics major at Rhodes College, stated<br />

that his experience at CLASSE was beneficial and rewarding; it helped convince him to pursue a doctoral degree instead of a<br />

bachelors or masters degree. For many REU students, the time spent in the program and on campus has encouraged them to<br />

seriously consider applying for graduate school at <strong>Cornell</strong> <strong>University</strong>. Seven students from past REU programs at Wilson Lab<br />

have continued their graduate work at <strong>Cornell</strong> and have enrolled in PhD programs or have pursued STEM (Science, Technology,<br />

Engineering, and Mathematics) teaching careers.<br />

Having a particle accelerator facility in the neighborhood is not only a unique attribute of the Ithaca community (how many<br />

cities advertise tours of a particle accelerator in their Visitor’s Guide), but it also offers remarkable opportunities for people<br />

throughout central New York to learn more about physics and other fields of science. Through the outreach program at CLASSE<br />

and nearly a dozen other STEM Outreach programs on campus, <strong>Cornell</strong> provides informal science education opportunities<br />

through a number of different of venues. A glimpse into the outreach program at the lab reveals staff involved in the following<br />

events: hosting informative, hands-on activity booths at local and state-wide science fairs; funding and offering workshops<br />

for campus-wide events such as Expanding Your Horizons and Career Explorations; participating in regional events to<br />

promote science literacy such as NanoDay and CCC Math Day; providing professional development training to educators from<br />

underperforming schools in New York City through the Science Sampler Series and Science Leadership Academy; escorting<br />

groups of community members through the tunnel that houses one of the most powerful tools for studying the atomic world;<br />

and many other examples. Through these volunteer efforts, and through the hard work of the outreach staff at CLASSE,<br />

thousands of people throughout the state of New York have been exposed to the many wonders of science. And hopefully,<br />

these people and others will recognize the important role that science research plays in their own lives and the lives of their<br />

children’s children.<br />

So, if you happen to be walking by my office<br />

and quickly glance at the vast collection of<br />

materials stacked on the shelves, see me<br />

carting tubs full of science equipment to and<br />

from the parking lot, or you see me escorting<br />

a young group of visitors throughout the<br />

Lab, ask me what I am doing and how you<br />

can help. The secret is out: I have the best<br />

job in the entire lab. And I will be more than<br />

delighted to share it with you!<br />

At the New York State Fair, Lora Hine explains to a visitor how<br />

potatoes create an electrochemical reaction between two probes<br />

in a Two Potato Clock which powers the digital clock. Visitors are<br />

invited to test various fruits and vegetables for themselves.<br />

Page 24 CHESS News Magazine 2009


MacCHESS Advances:<br />

Microcrystallography and More<br />

Marian Szebenyi<br />

Macromolecular Diffraction Division of<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Microcrystallography<br />

Crystals of proteins and other biological macromolecules<br />

never grow very big, due to the “soft” nature of the<br />

molecules and the weakness of the forces holding them<br />

together. At CHESS and other synchrotron sources, it is<br />

routine to deal with crystals that are ~100 μm across, but<br />

how often do we need to work with even smaller crystals<br />

Ithaca <strong>High</strong> School student Sarah Kim, under the direction<br />

of MacCHESS staffer Richard Gillilan, surveyed thousands of<br />

crystallization attempts (Fig. 1), and concluded that about<br />

50% of initial crystals were in the “micro” category,<br />

and that many larger crystals have “good”<br />

and “bad” regions and would benefit<br />

from use of a microbeam to isolate<br />

the better regions.<br />

• Mike Cook, Scott Smith, and Bill Miller have installed<br />

rear-mounted, piezo-based motors for crystal centering<br />

at F2. Using these, the runout error of spindle rotation<br />

can be reduced below 1 μm, as verified using a LabView<br />

program developed by <strong>Cornell</strong> student Ivan Temnykh<br />

(Fig. 3).<br />

Fig. 1: Sarah Kim and<br />

the crystal-scoring<br />

Java interface that<br />

she designed almost<br />

completely by herself,<br />

with some help from<br />

R. Gillilan.<br />

MacCHESS provides users with an x-ray beam as small<br />

as 5 μm, using a focusing capillary (see p. 63), excellent<br />

crystal-viewing optics to center a microcrystal in the beam,<br />

and a very precise rotation stage to keep it there. Users<br />

have determined important<br />

structures from “difficult”<br />

crystals using this equipment,<br />

e. g. the SAM riboswitch<br />

reported by Ailong Ke’s group 1<br />

(Fig. 2). Seeking to enhance the<br />

microcrystallography facility,<br />

• Don Bilderback’s group is<br />

working to reduce slope errors,<br />

and hence improve focusing, of<br />

X-ray capillaries.<br />

Fig. 2: The SAM riboswitch<br />

structure, rendered using Jmol 2 .<br />

Fig. 3: Piezomotor-equipped rotation stage<br />

installed in F2, Ivan Temnykh, plot of runout<br />

error using the modified stage.<br />

• Dave Schuller has upgraded the crystal-centering<br />

software to be more robust, and to include computer<br />

control of crystal illumination, as well as auto-centering<br />

using the XREC software 3 (Fig. 4).<br />

• Marian Szebenyi has added to the ADX data collection<br />

software an option for automated scanning of a crystal to<br />

identify the best-diffracting regions.<br />

A new experiment using an old technique<br />

With the very intense beam that would be produced by<br />

future sources such as an ERL, very tiny samples could<br />

produce diffraction patterns, but radiation damage would<br />

severely limit the amount of useful data that could be<br />

obtained from each sample. An experiment at D1, led by<br />

graduate student Sterling Cornaby, investigated the use of<br />

Laue diffraction to record a few images from each of several<br />

crystals mounted together (Fig. 5). With the Laue technique,<br />

many diffraction data are recorded simultaneously, with no<br />

need to rotate the sample; hence it is ideal for obtaining as<br />

much data as possible in a very short time. The experiment<br />

demonstrated the feasibility of collecting Laue data from<br />

a group of crystals and obtaining a structure from them<br />

CHESS News Magazine 2009 Page 25


Fig. 4: Java-based crystal-centering user interface, incorporating<br />

excellent video quality, automated centering, computer control<br />

of camera and light settings, easy recording of images of crystal,<br />

singly, as a rotation series or as a video stream.<br />

shown that modifications to the standard technique may<br />

be useful, and have also furthered understanding of what<br />

is going on inside cryocooled crystals. Matt Warkentin,<br />

from Rob Thorne’s lab, demonstrated that a faster cooling<br />

rate, obtained by blowing away the cold gas layer above<br />

the surface of liquid nitrogen before plunging a crystal<br />

into the liquid, is often advantageous 5 (Fig. 6). A pressure<br />

cryocooling technique developed primarily by Chae Un<br />

Kim (first a graduate student of Gruner’s, now a MacCHESS<br />

scientist) can also improve the quality of cryocooled<br />

crystals; see pages 43 and 71 for recent interesting<br />

results using pressure-cooling. Warkentin and Kim have<br />

also considered cooling crystals in capillaries, and have<br />

looked at the behavior of cryocooled crystals when the<br />

temperature is varied; this is an active area of research that<br />

will enhance our understanding of what is going on inside<br />

cold crystals, and may lead to better data from crystals<br />

which do not take kindly to routine cryopreservation.<br />

(see p. 63 for more details); further studies are planned to<br />

ascertain the limits of the technique.<br />

Detectors<br />

As in the past, MacCHESS has served as a test bed for<br />

new x-ray detectors. The first of the latest generation<br />

of CCD-based detectors from Area Detector Systems<br />

Corp. (ADSC), the Q-270, was installed at CHESS in 2006<br />

for testing; a few problems (mostly in software) were<br />

discovered and corrected, and the detector was made<br />

available to users later that year. Generally similar to the<br />

earlier Q-210 and Q-315, the new detector differs in having<br />

increased sensitivity, for accurate measurement of weak<br />

reflections. ADSC is also collaborating with the Gruner<br />

group in developing a tiled pixel array detector (PAD) for<br />

crystallographic use. A prototype PAD was tested at CHESS<br />

in 2007, with promising results 4 . Modifications to ADX<br />

allowed easy switching between CCD and PAD devices,<br />

as well as collection of a series of PAD images at different<br />

detector locations, which could subsequently be<br />

assembled into a large image. The sensitivity of<br />

the PAD proved to be similar to that of the Q-210<br />

CCD, but the dynamic range is much higher<br />

and the readout time much less. No geometric<br />

distortion corrections are necessary for the<br />

PAD, and background subtraction is easily done<br />

with existing software; yet to be developed are<br />

corrections for edge effects, and procedures to<br />

scale and combine multiple small PAD images<br />

into one in real time. See p. 54 for the latest on<br />

PAD’s.<br />

Temperature and pressure<br />

Typically, macromolecular crystals are<br />

cryocooled by plunging into a container of liquid<br />

nitrogen, and then maintained at about 100 K<br />

for data collection. Recent experiments have<br />

Page 26 CHESS News Magazine 2009<br />

Fig. 5: Collection of microcrystals on a MicroMesh mount, for<br />

use in Laue experiment at D1.<br />

SAXS and SAD<br />

Over the last few years there has been increasing interest<br />

by structural biologists in conducting small angle x-ray<br />

scattering (SAXS) studies, and CHESS now offers facilities<br />

for SAXS including a new temperature-controlled sample<br />

cell, helium-filled flight paths, and expert<br />

advice. See the article on p. 50 for a full<br />

report on bioSAXS. The SAD (singlewavelength<br />

anomalous diffraction)<br />

technique, using the signal from heavy<br />

atoms such as Se or Hg, is the most<br />

common way to phase structures when<br />

no appropriate model is available for<br />

molecular replacement. At CHESS, A1 is<br />

suitable for Se-SAD and F1 for Br-SAD,<br />

and F2 can be tuned to accommodate<br />

a number of elements with absorption<br />

edges between 7 and 14 keV. For<br />

elements with edges at lower energies,<br />

Fig. 6: Simple “hyperquenching” device for<br />

blowing away cold gas above the surface of<br />

a liquid nitrogen bath.


down to slightly below 6 keV, the F3<br />

station, with multilayer optics, presents<br />

another option; see sidebar.<br />

Station upgrades<br />

A number of changes have been made to<br />

improve users’ productivity and comfort:<br />

Station F3<br />

Ulrich Englich<br />

Macromolecular Diffraction Division of<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

• Robust safety shields which open and<br />

close automatically.<br />

• Motorized detectors which roll back<br />

automatically when a hutch is opened,<br />

and which can be moved using ADX.<br />

• Up-to-date computing hardware<br />

and software, including an extensive<br />

collection of standard crystallographic<br />

software.<br />

References:<br />

1. C. Lu, A.M. Smith, R.T. Fuchs, F.<br />

Ding, K. Rajashankar, T.M. Henkin,<br />

and A. Ke; “Crystal Structures of the<br />

SAM-III/S(MK) Riboswitch Reveal the<br />

SAM-dependent Translation Inhibition<br />

Mechanism”, Nat. Struct. Mol. Biol. 15,<br />

1076-1083 (2008)<br />

2. “Jmol: an open-source Java viewer<br />

for chemical structures in 3D”, http://<br />

www.jmol.org<br />

3. “Crystal Recognition” , http://www.<br />

embl-hamburg.de/XREC<br />

4. W. Vernon, M. Allin, R. Hamlin, T.<br />

Hontz, D. Nguyen, F. Augustine, S.M.<br />

Gruner, Ng.H. Xuong, D.R. Schuette,<br />

M.W. Tate, and L.J. Koerner; “First<br />

Results from the 128x128 Pixel Mixedmode<br />

Si X-ray Detector Chip”, SPIE<br />

Optics & Photonics Conference, San<br />

Diego, CA (August 26-30, 2007)<br />

5. M. Warkentin, V. Berejnov,<br />

N.S. Husseini, and R.E. Thorne;<br />

”Hyperquenching for Protein<br />

Crystallography”, J. Appl. Cryst. 39,<br />

805-811 (2006)<br />

The F3 station is shared between CHESS (e.g. x-ray topography studies,<br />

x-ray imaging, x-ray fluorescence) and MacCHESS (macromolecular<br />

crystallography). Darren Dale, the responsible scientist for CHESS,<br />

maintains all beamline components up to the hutch and is currently<br />

working on further enhancements. The station supports a variety of<br />

monochromators: Si or Ge crystals as well as multilayers of various<br />

bandwidths. For macromolecular experiments the new 0.22% multilayers<br />

are used. Recent improvements in the monochromator box now allow<br />

users to tune the energy between 5.9 and 15.4 keV (2.1Å – 0.81 Å). This<br />

range spans the K and L edges of most commonly used elements for<br />

anomalous dispersion experiments.<br />

In the hutch CHESS and MacCHESS have rolling setups on separate<br />

optical tables for quick and easy switch-over. A new 1.5m optical<br />

table has been designed for crystallographic purposes; it supports all<br />

experimental components on one platform. This greatly facilitates<br />

alignment procedures and overall stability. It also gave CHESS the<br />

opportunity to install a new “integrated motion controller and drivers<br />

unit” for the table motors, which is much more compact than the old<br />

controller. In addition, the 32 bit workstation for data collection and<br />

processing has been replaced with a modern 64 bit Linux based system<br />

similar to those at other MacCHESS beamlines.<br />

The station has been used mainly by MacCHESS personnel, the Crane<br />

group from <strong>Cornell</strong> and the Cingolani group from Upstate Medical<br />

<strong>University</strong>. Recent experiments at the station examined samples that<br />

contain anomalous scatterers, e. g. Xe, Fe and S. The results show that<br />

we can obtain an anomalous signal from an element as weak as sulfur,<br />

with data collected at 7.1 keV. Due to the increase in flux with the 0.22<br />

% multilayer monochromator, data collection times were significantly<br />

shorter (~ 5 times) compared to a data set taken from a similar sample at<br />

the F2 beamline. As an example of the quality of anomalous data from<br />

F3, the figure shows an anomalous difference Fourier map from horse<br />

hemoglobin, after refinement, contoured at 4 σ. The region depicted<br />

includes both heme groups and their protein surroundings. Large<br />

positive peaks (blue) appear at the positions of the iron atoms in the<br />

center of the hemes, and smaller peaks are visible for sulfurs, e. g. in the<br />

disulfide bond located slightly above and to the right of the left-hand Fe.<br />

CHESS News Magazine 2009 Page 27


XPaXS: Improved Data Acquisition<br />

and Real-Time Analysis at CHESS<br />

The Scanning X-ray Fluorescence Microscopy (SFXM) capability at CHESS draws users from diverse fields, including Environmental<br />

Science, Soil Science, Anthropology and Native American Studies, Dendrochronology and Classics. When I inherited responsibility<br />

for the F3 beamline, two of the elements essential for the success of this research program were already available: wide-bandpass<br />

multilayers to provide the intense monochromatic flux required for detection of elements in trace quantities; and singlebounce<br />

capillary x-ray optics capable of yielding microbeams at the sample position. In the past three years we have completely<br />

refurbished the F3 station and upgraded all of the sample manipulation hardware, which is now capable of scanning areas 30 cm<br />

wide by 30 cm tall and 5 cm deep, with step sizes below 1 micrometer if necessary.<br />

Scanning large areas with high resolution presents some new experimental challenges. To give a sense of scale, scanning an area 3<br />

mm by 3 mm with a 15 µm step size requires 40,000 independent measurements, and if integrated for 1 second per point the scan<br />

will take more than 11 hours to finish (development of new multi-element<br />

detectors will improve the throughput in the near future). We use the “spec”<br />

command-line program for data acquisition, which is familiar to seasoned<br />

synchrotron researchers but frustrating to many newcomers, see figure 1.<br />

When I first took over F3, we would use “spec” to start a scan, and then watch<br />

raw data scroll across the screen for 11 hours. The experience was similar to<br />

watching the screens of encrypted data in the movie “The Matrix”, where a<br />

willful suspension of disbelief is required to accept that the data are somehow<br />

meaningful.<br />

Many experiments required recording the full fluorescence spectrum at each<br />

point. The datasets were growing in size and becoming difficult to work with.<br />

Some peak-fitting routines were available to process the data after collection<br />

was complete, but on several occasions we discovered problems with a<br />

Fig. 1: Everybody’s first experience with the command line.<br />

dataset weeks after the experiment had ended. Clearly what we needed was<br />

an improved analysis capability, one that could fit fluorescence spectra with many overlapping peaks, determine the elemental<br />

concentrations from the peak areas, and update an element map in a graphical user interface in real time while the experiment<br />

is ongoing. We also wanted the ability to interact with the data in real time without disrupting the acquisition, for example by<br />

selecting a single pixel or region of interest in the image map and inspecting the raw fluorescence spectra and fit results or<br />

performing some kind of statistical analysis such as Principle Components Analysis or cluster analysis.<br />

In order to foster development of the multidisciplinary research program at F3, we have begun augmenting the flexible<br />

command-line interface with just such an intuitive graphical user interface for experimental controls and data processing (see<br />

figure 2). Working with <strong>Cornell</strong> undergraduate students, we have begun to develop XPaXS, an extensible set of packages for x-ray<br />

science. These packages are written in Python, which is an established, powerful cross-platform scripting language that allows<br />

scientists to focus more on science than<br />

the minutiae of computer programming.<br />

Python is emerging as the standard language<br />

for systems integration, in part due to its<br />

permissive open-source license (it can be<br />

used for both free and commercial projects),<br />

simple and elegant syntax and wide variety<br />

of scientific libraries, all of which enable<br />

rapid development of powerful, large and<br />

yet maintainable projects. Our software<br />

development work is done in cooperation<br />

with other successful free software projects.<br />

Page 28 CHESS News Magazine 2009<br />

Darren Dale<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Fig. 2: Graphical user interface for HDF5 files<br />

and spec scan controls.


For example, SpecClient provides a<br />

python interface to the “spec” data<br />

acquisition program, and PyMca is<br />

used for advanced spectral fitting and<br />

composition analysis, both of which were<br />

developed at the European <strong>Synchrotron</strong><br />

Radiation Facility.<br />

The graphical user interface and<br />

improved data handling allow us to<br />

develop flexible and powerful tools for<br />

data inspection and analysis. Experiments<br />

that once provided only lines of text on<br />

the command line as feedback during the<br />

measurement now provide capabilities<br />

for real-time analysis and interactive<br />

inspection of data in a form that is most<br />

meaningful to the experimenter. The<br />

program is multithreaded, so it is possible<br />

to interact with the data even while both<br />

data acquisition and analysis are ongoing.<br />

It is possible to select a single point in<br />

a map, or drag out a region of interest,<br />

and inspect the fluorescence spectra to<br />

identify anomalies or bad data points, or<br />

to average over a heterogeneous sample<br />

in order to calibrate the measurement<br />

to a NIST standard. In the example<br />

illustrated in figure 3, all the spectra<br />

in a region of interest specified in the<br />

calcium mass fraction map are passed<br />

to PyMca for peak fitting<br />

and concentration<br />

determination, and<br />

the averaged results<br />

are compared with the<br />

accepted values.<br />

This acquisition and<br />

analysis package is<br />

developed with the hope<br />

that many will find it<br />

useful, and that many<br />

will contribute to — and<br />

take credit for — its<br />

development. Working<br />

with colleagues at the<br />

European <strong>Synchrotron</strong><br />

Radiation Facility, we have<br />

settled upon a simple<br />

common file format based<br />

upon the freely available<br />

HDF5 technology suite,<br />

which supports the<br />

compressed, binary<br />

HDF5 format developed<br />

specifically for scientific<br />

applications. HDF5 allows<br />

us to efficiently work<br />

with arbitrarily large and<br />

complicated datasets,<br />

loading only small subsets of the data<br />

into memory at any given time. We take<br />

advantage of computers with multiple<br />

processors and even computing clusters<br />

to process data in parallel, as quickly as<br />

possible. Many synchrotron techniques<br />

can be significantly enhanced by the<br />

ability to analyze and inspect relatively<br />

large datasets in various ways, both in<br />

real time during the acquisition and<br />

off-line after acquisition is complete.<br />

The software is designed to be easily<br />

extended and applied to other kinds of<br />

experiments and analyses, has already<br />

been used for analysis of EXAFS data,<br />

and is currently being extended to other<br />

kinds of experiments at CHESS as well.<br />

Users can revisit or reprocess their data<br />

after leaving the lab using the Python<br />

packages which are freely available on<br />

the internet. XPaXS and its dependencies<br />

are free, open-source projects that can be<br />

used on Linux, Windows and Macintosh<br />

OS-X. The source code is also available for<br />

download, so anyone can add whatever<br />

additional features they need, regardless<br />

of platform. Everyone is encouraged<br />

to contribute and take credit for their<br />

contributions.<br />

Documentation is a major focus of our<br />

project. A Users’ Guide is currently being<br />

developed, along with a Developers’<br />

Guide and Application Programming<br />

Interface reference to help those<br />

interested in adding new features.<br />

The documentation is created using<br />

modern documentation generation<br />

tools. The library documentation, online<br />

html and printable pdf documents<br />

are all generated from a common<br />

origin, so the entire documentation<br />

suite can easily be kept up to date.<br />

The documents are searchable, crossreferenced,<br />

and automatically indexed.<br />

The documentation in its current form<br />

is available at http://www.chess.cornell.<br />

edu/software/xpaxs.<br />

The documentation is a central and<br />

ongoing effort. The users guide for<br />

the SXFM program will be the primary<br />

focus in the near future, in order to<br />

provide users who wish to work more<br />

independently with the means of<br />

doing so. A comprehensive set of unit<br />

tests will be assembled to aid in future<br />

development and ensure trouble-free<br />

operation. We are developing several new<br />

features, such as a network-based video<br />

monitoring capability which is currently<br />

being developed to allow<br />

remote observation of the<br />

experiment and its progress.<br />

We are also planning to<br />

incorporate existing code<br />

with features similar to<br />

those available at the CXRO<br />

website for calculations<br />

of x-ray transmission,<br />

anomalous form factors,<br />

index of refraction, etc.<br />

The long range goal is to<br />

provide a framework such<br />

that additional packages<br />

can be added over time to<br />

provide all the components<br />

necessary for imaging,<br />

spectroscopy and scattering<br />

experiments, optics design<br />

and characterization,<br />

characterization of x-ray<br />

sources, and most likely<br />

other areas we have not yet<br />

considered.<br />

Fig. 3: (Top) Calcium mass fracton<br />

distribution in NIST citrus leaves<br />

standard reference. (Bottom) Spectrum<br />

averaged over a region of interest<br />

selected from the Calcium map and the<br />

fit produced by PyMca.<br />

CHESS News Magazine 2009 Page 29


CHESS <strong>High</strong>-pressure Stations (B1/B2) Update<br />

Zhongwu Wang<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

<strong>High</strong> pressure coupled with synchrotron x-ray diffraction is one powerful technique for discovering novel<br />

physics and chemistry with a wide spectrum of applications. CHESS has long been playing a significant role<br />

in the development of high pressure synchrotron facilities. To expand our research abilities, particularly in<br />

nanoscience and nanotechnology, we have made an effort to improve and update our high pressure facilities,<br />

to be capable of a series of new in-situ measurements of materials, including wide and small angle x-ray<br />

diffraction and Raman measurements at extreme conditions of pressure and temperature. Here, we highlight<br />

several modifications and developments at CHESS:<br />

1.<br />

2.<br />

3.<br />

4.<br />

Two beam lines for energy-dispersive and angle-dispersive x-ray diffraction are combined and<br />

merged into one versatile station. This modification not only keeps all experimental capabilities<br />

- energy-dispersive or angle-dispersive x-ray diffraction can be selected by repositioning two<br />

monochromator crystals - but also expands the hutch space, enabling introduction of a series of<br />

fixed and portable instruments for development of new techniques. The improved capabilities<br />

include a three-dimensional rotating stage for side x-ray diffraction, laser-excited Raman<br />

spectroscopy, a detector control system, a high temperature tuning system etc.<br />

Four-dimensional control for movement of the Mar345 detector is implemented. Slightly tilting and<br />

rotating the Mar345 detector enables an accurate calibration of the sample-to-detector distance<br />

and associated parameters. Most importantly, a new in-situ technique has been developed allowing<br />

measurement of both small and wide angle x-ray diffraction at various pressures and temperatures<br />

without multiple calibrations. Easy optimization of x-ray energy and sample-to-detector distance<br />

makes the two types of measurements possible during one single pressure run.<br />

Switching the x-ray source to white beam and using a large area image plate detector enable single<br />

crystal Laue diffraction under pressure.<br />

A series of side instruments become operational for users. These include two portable<br />

spectrometers for Raman, reflectivity and transmission spectra, mechanical and Electrical Discharge<br />

Machining (EDM) drillings, a high pressure gas loading system, optical microscopes, etc.<br />

Page 30 CHESS News Magazine 2009


The <strong>Cornell</strong> <strong>Energy</strong> Recovery Linac:<br />

update on a future source of coherent hard x-rays *<br />

Don Bilderback 1,2 , Bruce Dunham 3 , Georg Hoffstaetter 3 ,<br />

Alexander Temnykh 3 , and Sol Gruner 1,4<br />

1<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

School of Applied Physics, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

Laboratory of Elementary-Particle Physics, <strong>Cornell</strong> <strong>University</strong><br />

4<br />

Physics Department, <strong>Cornell</strong> <strong>University</strong><br />

*<br />

For the <strong>Cornell</strong> ERL Team, <strong>Cornell</strong> <strong>University</strong><br />

<strong>Synchrotron</strong> radiation research is still growing, with newer and more capable x-ray sources being constructed. Petra III<br />

in Hamburg, Germany and the NSLS II in Upton, NY are examples of storage rings under construction that are nearing<br />

the expected limits of storage ring performance. In contrast, linac-based sources such as energy recovery linacs (ERLs)<br />

and x-ray free electron lasers (XFELs) are still at an early stage of development, and have great potential for steady<br />

improvements in performance for many years to come.<br />

The ERL upgrade to the CESR ring described below<br />

promises to generate ultra-bright electron beams, and<br />

thus ultra-bright x-ray beams with dramatically smaller<br />

emittances and shorter pulse durations than those available<br />

from storage rings 1 . The high coherence and temporal<br />

properties of the ERL will have transformative, broad<br />

impact across the sciences and engineering by enabling<br />

numerous experiments that are now not feasible at even<br />

the most modern 3 rd generation storage ring sources;<br />

nor would many of these experiments be suitable for<br />

x-ray free electron lasers. Enabled science includes many<br />

types of imaging of materials and biological structures,<br />

macromolecular structure determination without<br />

crystals, effective nanoprobes for matter at higher static<br />

pressures than are presently feasible, and detailed studies<br />

of the structure of glasses, disordered materials, and<br />

polycrystalline materials. The ERL would be especially<br />

useful in cases where structure needs to be determined on<br />

objects for which structural details vary from one specimen<br />

to another, which is the case for most nanoparticles and<br />

biological cells and organelles. It would also allow the<br />

recording of extremely rapid events, so researchers<br />

can “visualize” rapid physical, catalytic and<br />

biological processes taking place on times scales<br />

down to 50 femtoseconds. Societal impacts<br />

will come in numerous areas such as life-saving<br />

medicines and drugs, new materials for better<br />

batteries, more efficient catalysts for fuel cells,<br />

better composite materials, smarter electronics, etc.<br />

The transverse emittance of the electron bunches<br />

used to generate x-rays determines the spectral<br />

brilliance and transverse coherence of the x-ray<br />

beams. Ideally, the emittance should be small<br />

enough to produce full transverse coherence of the<br />

x-rays, i.e., to produce a diffraction-limited x-ray<br />

beam. The underlying basis of an ERL is that, in<br />

principle, very low emittance electron beams can<br />

be generated from a laser-driven photocathode,<br />

and accelerated to high (GeV) energies without<br />

substantial emittance growth in linear accelerators.<br />

<strong>High</strong> brightness, highly coherent x-ray beams can then<br />

be generated from these electrons as they pass through<br />

undulators. Since the electron beam carries several<br />

hundred megawatts of beam power, this is feasible only<br />

if the electron beam energy is recovered after the x-rays<br />

are generated. Using a superconducting (SC) RF linac,<br />

essentially all the electron energy may be recovered by<br />

passing the beam through the linac a second time, 180 o<br />

out of phase with the accelerated beam, hence, the name<br />

<strong>Energy</strong> Recovery Linac. Although the basis of the ERL idea<br />

was suggested many years ago by Maury Tigner 2 , it only<br />

became practical for x-ray generation in the mid-1990’s,<br />

due to advances in superconducting linac technology and<br />

photoemission electron sources 3 .<br />

Phase 1a & 1b: Testing the first stages of the ERL accelerator<br />

The ERL group at CLASSE is in the midst of prototyping<br />

the technology that will be needed for a 5 GeV ERL facility.<br />

The most difficult and critical component in need of R&D<br />

is the injector; a full-scale prototype injector has now been<br />

assembled in the L0 area of Wilson Laboratory (Figure 1).<br />

The injector generates a beam of high average current, lowemittance<br />

electrons using a laser-driven photo-cathode and<br />

then accelerates them to relativistic energies up to 15 MeV.<br />

Fig. 1: Floor layout of the ERL test area in L0 of Wilson Laboratory, just south<br />

of CHESS East. From right to left: DC photocathode, superconducting linac,<br />

branch lines for diagnostics, and finally, a high power beam dump.<br />

CHESS News Magazine 2009 Page 31


A consequence of special relativity is that the mutual charge repulsion that tends to blow apart a bunch of electrons at low<br />

energy is counteracted at relativistic speeds. Thus, the difficult task of the injector is not only to generate a low emittance<br />

electron beam, but to preserve the emittance until relativistic energies are achieved. Figure 1 shows the layout of the test<br />

accelerator designed to generate ultra-low emittance bunches of electrons at a 1.3 GHz rate with up to 100 mA average<br />

current. This injector is now in the commissioning stage.<br />

The electrons are produced in a special cathode material using the photoelectric effect. A precisely shaped laser pulse is<br />

directed at the cathode, giving the electrons enough energy to escape. The ejected electrons are then accelerated across<br />

a small gap using a high voltage power supply. The cathode material is gallium arsenide (GaAs), which is commonly used<br />

for high speed electronics, in photomultiplier tubes, and in night vision goggles. To obtain efficient electron emission, the<br />

surface of the GaAs wafer must be as clean as possible and maintained at a vacuum level below 10 -11 torr. After cleaning,<br />

Fig. 2: Photocathode gun area. A vacuum of better than 10 -11 Torr<br />

has to be maintained around the photocathode to avoid ion back<br />

bombardment of the photocathode, which would spoil the lifetime of<br />

the photocathode by ruining the monolayer of Cs on the surface. A<br />

load-lock chamber facilitates quick change of the cathodes for testing<br />

and for renewing the Cs monolayer needed for the cathode to work well.<br />

Fig. 3: The superconducting (SC) linac consists of 5 SC cells<br />

in series. Each cell uses a waveguide to conduct 1.3 GHz<br />

microwave power from associated klystron transmitters on the<br />

mezzanine floor above. The cavities are cooled by liquid helium<br />

at 1.8 K from a closed-cycle refrigeration system. The linac is<br />

now functional.<br />

it is coated with very thin layers of cesium and fluorine, which act to reduce the work function, making it easier for the<br />

electrons to escape (Figure 2 shows the cathode preparation system). The properties of GaAs have been extensively<br />

measured 4,5 , demonstrating that it is currently the best cathode choice for obtaining low beam emittance and high average<br />

current.<br />

A very sophisticated laser system is required to produce bunches of electrons with the correct shape, power and time structure. It<br />

turns out that the shape of the outgoing electron pulse closely matches the shape of the incoming laser pulse, and the electron<br />

beam emittance can be reduced by choosing the correct shape. The spacing (frequency) of the pulses must also match the<br />

frequency of the RF accelerating cavities, which operate at 1.3 GHz. There are no commercial laser systems available to meet all of<br />

the needs for an ERL injector, thus we have built a custom laser using ytterbium (Yb) doped optical fibers.<br />

After the electron bunches are produced<br />

at the cathode and given an initial<br />

acceleration by the electron gun, they are<br />

boosted to high energy using a series of<br />

5 superconducting RF cavities (Figure 3).<br />

Each cavity can transfer up to 100 kW of<br />

microwave energy to the electron beam, for<br />

a total of 500 kW (1/2 of a million watts!).<br />

Superconducting cavities are used to<br />

overcome the heating problems that would<br />

occur with such large power. Figure 4<br />

shows the five high-power microwave tubes<br />

that provide the energy to the accelerating<br />

cavities.<br />

Fig. 4: Microwave generators<br />

produce the power needed to<br />

drive the superconducting linac.<br />

Page 32 CHESS News Magazine 2009


The final parts of the test set-up are diagnostic tools to measure the emittance of the electron bunches and its pulse duration<br />

before the beam is sent to a 0.5 MW beam dump.<br />

We are now starting the second phase of the ERL R&D program. The proposed research program addresses five critically important<br />

technology areas:<br />

1. Superconducting linac development. The ERL requires linacs capable of continuous duty (CW) operation at high current with<br />

short bunch lengths while preserving emittance and dealing effectively with higher order modes. A full linac cryomodule<br />

(somewhat similar to the injector linac, but with longer cavities and with much weaker power couplers because in the<br />

ERL it is the decelerated beam that fills the cavity with energy, not the power coupler) will be designed and fabricated at<br />

<strong>Cornell</strong>. Cryomodule component testing will be done collaboratively with several other institutions. The program includes<br />

development of requisite cavities, higher-order-mode absorbers, beam position monitors, in-vacuum cryomagnets, RF<br />

amplifiers and power couplers, and cool-down procedures.<br />

2. <strong>High</strong> brightness photoinjector development. Continued development will be performed on laser driven photoinjectors capable<br />

of delivering requisite current at high brightness and long photocathode lifetime. This includes R&D on new photocathodes,<br />

high voltage electron sources, drive lasers, and advanced beam simulation tools. Further R&D on the HV insulator is needed<br />

to reach the required 500 to 750 kV energy level from the electron gun.<br />

3. <strong>High</strong>-brightness electron beam physics and diagnostics. The program will address challenges of emittance preservation from<br />

the electron source, through the injector, merger, and transport loop sections via experiments and simulations on the <strong>Cornell</strong><br />

injector. Beam diagnostics to monitor bunch timing and position will be advanced, both for beam stability at the x-ray source<br />

points, and for simultaneously accelerating and decelerating bunches in the linacs.<br />

4. Beam dynamic effects will be investigated, such as ion trapping, short bunch behavior and other possible sources of instability.<br />

Beam loss mechanisms, impedance issues, failure modes and beam abort strategies will all be studied through overall system<br />

modeling. In each case, methods of mitigation and control will be studied.<br />

5. X-ray beamline components necessary to exploit ERL capabilities will be developed. These include novel undulators to exploit<br />

the identical horizontal and vertical emittances of an ERL, x-ray optics for high specific heat loads, coherence monitors, and<br />

detectors to effectively utilize the temporal possibilities of a coherent ERL source.<br />

Fig 5: Low-energy electrons are produced and accelerated to 5-10 MeV at the injector. They are raised to 2.5 GeV in the North Linac by traveling “on-crest”<br />

through the RF fields produced by the SRF cavities, turned around and raised to 5 GeV after passing through the South Linac. X-rays will be produced in<br />

undulators in the position of the blue arrows, mostly in a new building to be added to the east of Wilson laboratory. Electrons returning to the North Linac<br />

are shifted 180 degrees out of accelerating phase (by adjusting the path length) so that the electrons are now “off-crest” and their energy is recovered in<br />

both linacs before the spent electron beam is sent to the dump. There is room in the layout for a non-energy-recovered, fast-pulse beamline as shown by<br />

the red arrow.<br />

Phase II: the full-energy, 5 GeV machine<br />

The work of the Phase I development (described above) is designed to provide confidence and costing for a 5 GeV ERL that we<br />

would hope to build at Wilson Laboratory.<br />

The CESR tunnel and Wilson laboratory infrastructure will be reused for the ERL facility. Much of the ERL would consist of new<br />

construction to the East (above, in Figure 5) of the present CESR and Wilson Lab complex. A new partially-underground laboratory<br />

will be added to increase the number of x-ray beamlines and new tunnels will be bored to house the new linacs required. We<br />

already have a first round of plans and costing from ARUP, an internationally-based architectural and engineering company<br />

that has been contracted to develop plans for the civil engineering (buildings and tunnels) of the project. The present plans are<br />

CHESS News Magazine 2009 Page 33


for conceptual design purposes. We are presently optimizing the layout of tunnels, buildings and the machine to fit campus,<br />

performance, and cost constraints. For example, we are evaluating layouts with twin linacs in the same tunnel as well as versions<br />

in separate tunnels. When this work is completed, we plan to submit a conceptual design proposal to the NSF for a full-scale<br />

5 GeV ERL facility, including x-ray beamlines.<br />

The ERL design group is working on a<br />

Conceptual Design Report containing the<br />

following parameters table which shows<br />

various ERL operating modes. An ERL is<br />

more flexible than a storage ring, which<br />

opens new opportunities and a plethora<br />

of operating modes.<br />

There is a long list of items that are being<br />

worked on in parallel with work on the<br />

injector and the 5 GeV facility layout,<br />

including electron beam machine design,<br />

shielding wall design, collimators ahead<br />

of undulators, exit crotch design, etc.<br />

The x-ray group has begun planning for<br />

5 to 25 m long undulators and beamlines<br />

that extend outward to 70 m from the<br />

source. We have established preliminary<br />

working groups to conceive of beamlines<br />

in the following areas:<br />

1. Diffraction-limited x-ray scattering<br />

beamline suitable for soft-matter<br />

using SAXS, USAXS, WAXS and<br />

GISAXS techniques;<br />

2. Short-pulse repetitive pump-probe timing beamline for 100 femtosecond pulse work at a repetition rate of 100 kHz and a 2<br />

picosecond pulse length at a 1.3 GHz repetition rate. This will be useful for ultrafast time-resolved experiments;<br />

3. A beamline for coherent diffractive imaging and dynamic studies of bulk materials, interfaces and biological samples<br />

including techniques of intensity fluctuation spectroscopy, ptychography. Examples of application areas are dynamics of<br />

magnetic materials and the glass transition;<br />

4. Inelastic x-ray scattering (IXS) with meV resolution will be used to study the dynamical properties of materials at atomic to<br />

mesoscopic length scales. An ERL IXS facility has the potential to outperform all existing facilities, their proposed upgrades,<br />

and sources under construction by one order of magnitude more flux;<br />

rd<br />

5. A nanoprobe for making a 1 nm diameter x-ray beam at 10 keV with as much flux per square nm as most 3 generation light<br />

sources put into a square micron. This will make possible x-ray experiments on even a single atom, which might be useful<br />

for studying the in-situ doping properties (atom clusters that are electrically inactive) of the smallest line-width transistors or<br />

performing x-ray experiments in the smallest possible volumes.<br />

New undulator designs are underway:<br />

ERLs offer opportunities to use insertion devices<br />

that take advantage of long, small, round bores (5<br />

mm). This is very different than in a storage ring,<br />

where a much wider horizontal bore is needed to<br />

provide orbital room for electron injection. The small<br />

sized bore feasible with an ERL allows design of very<br />

compact magnetic structures that can generate<br />

larger magnetic fields than the planar undulators<br />

used in 3 rd generation storage rings.<br />

Figure 6 shows a novel “Delta” undulator that was<br />

designed to utilize the unique properties of ERL<br />

electron beams 6 . As opposed to conventional<br />

undulators, it provides full control of x-ray<br />

polarization and a stronger magnetic field that<br />

Page 34 CHESS News Magazine 2009<br />

Fig. 6: Prototype of<br />

the Delta undulator<br />

made from small<br />

triangular blocks of<br />

NdFeB material with<br />

a remnant field of<br />

1.26 Tesla. The central<br />

hole for the electron<br />

beam is only 5 mm in<br />

diameter.


translates to high x-ray flux. In addition, it is<br />

much more compact and cost-efficient. To<br />

demonstrate the new design principles, a 30<br />

cm long model has been built and recently<br />

tested with a Hall probe, Figure 7.<br />

Testing of the very first Delta-type undulator<br />

prototype revealed a 1.25T peak field in planar<br />

mode and 0.85T in helical (Figure 8) that is<br />

approximately 90% of the design value.<br />

The field profile analysis indicates that the field<br />

errors cause less than a 3% loss of the x-ray<br />

photon flux in the helical mode and 20% loss<br />

in the planar mode. These very favorable test<br />

results have confirmed the basic principles of<br />

the Delta-style design.<br />

Fig. 7: Delta undulator model<br />

on magnetic field measurement<br />

bench, complete with coils for<br />

compensating for background<br />

magnetic fields. The center<br />

ceramic tube guides the Hall<br />

probe down the bore of the<br />

undulator during the magnetic<br />

profile measurement.<br />

Fig. 8: The measured<br />

magnetic field components<br />

in planar and in helical mode<br />

of operation corresponding<br />

to linear and circular x-ray<br />

polarization. Plots are from<br />

reference [7].<br />

Presently we are working on the next, fully-functioning model. It will have an improved mechanical design which should increase<br />

peak field up to the full design value and reduce magnetic field errors still further. The next model will be UHV compatible and will<br />

be tested with a small diameter electron beam at a linac test facility.<br />

Conclusion:<br />

A 5 GeV ERL facility would be a great advance to synchrotron radiation science and would have numerous benefits to society. R&D<br />

on necessary ERL components is well underway. Plans for a full-scale, 5 GeV hard x-ray ERL are in an advanced stage of planning.<br />

We intend to deliver a conceptual design for the full-energy ERL within the next year.<br />

References:<br />

1. C. Sinclair, and S. Gruner; “ERL Developments”, CHESS News Magazine, pg 9 (2005)<br />

2. M. Tigner; Nuovo Cimento 37, 1228 (1965)<br />

3. B. Aune et al.; Phys. Rev. ST-AB 3, 092001 (2000); G. R. Neil et al., Phys. Rev. Lett. 84, 662 (2000); C. K. Sinclair,<br />

Proceedings of the 2003 Particle Accelerator Conference, p. 76<br />

4. I. Bazarov et al.; “Thermal Emittance and Response Time Measurements of NEA Photocathodes”, J. Appl. Phys. 103, 05490<br />

(2008)<br />

5. I. Bazarov, B. Dunham, and C. Sinclair; “Maximum Achievable Brightness from Photoinjectors:, Phys. Rev. Lett. 102,<br />

104801 (2009)<br />

6. A. Temnykh; “Delta Undulator for <strong>Cornell</strong> <strong>Energy</strong> Recovery Linac”, Phys. Rev. ST Accel. Beams 11, 120702 (2008)<br />

URL: http://link.aps.org/doi/10.1103/PhysRevSTAB.11.120702<br />

7. A. Temnykh; “Evaluation of Magnetic and Mechanical Properties of Delta Undulator Model”, Reprint CBN 09-1,<strong>Cornell</strong><br />

(2009) URL: http://www.lns.cornell.edu/public/CBN/2009/<br />

CHESS News Magazine 2009 Page 35


New Mechanical Testing Methods for Structural Materials at Small Size Scales<br />

Matthew Miller<br />

Alteration of Citrine Structure by Hydrostatic Pressure<br />

Sol Gruner, Buz Barstow, Nozomi Ando, and Chae Un Kim<br />

Putting Color into Surface Diffraction<br />

Detlef Smilgies<br />

Self-Assembled Nano-Checkered Thin Films Studied by Reciprocal Space Mapping at<br />

CHESS<br />

Sean O'Malley, Peter Bonanno, Keun hyuk Ahn, Andrei Sirenko, Alex Kazimirov, Soon-Yong Park, Yoichi Horibe,<br />

and Sang-Wook Cheong<br />

Structures from Solutions: Biomolecular Small-Angle Solution Scattering at MacCHESS<br />

Richard Gillilan<br />

Development of X-ray Pixel Array Detectors<br />

Sol Gruner<br />

Exploring New Physics of Nanoparticle Supercrystals by <strong>High</strong> Pressure Small Angle<br />

X-ray Diffraction<br />

Zhongwu Wang, Ken Finkelstein, and Detlef Smilgies<br />

Topography of Diamonds at CHESS Helps Nuclear Physics Program at JLAB<br />

Richard Jones, Franz Klein, and Ken Finkelstein<br />

Nanoparticle-block Copolymer<br />

Ulrich Wiesner, Laura Houghton, and Sol Gruner<br />

Heat-bump Measurements at CHESS A2 Wiggler Beam<br />

Peter Revesz, Alex Kazimirov, Ivan Bazarov, Jim Savino, Emmett Windisch, and Christopher MacGahan<br />

Advances in X-ray Microfocusing with Monocapillary Optics at CHESS<br />

Sterling Cornaby, Thomas Szebenyi, Heung-Soo Lee, and Don Bilderback<br />

Carbon Nanotube and Crystalline Silicon Structures<br />

Eric Meshot, Mostafa Bedewy, Sameh Tawfick, K. Anne Juggernauth, Eric Verploegen, Yongyi Zhang, Michael De Volder,<br />

and John Hart<br />

Phase Behavior of Water inside Protein Crystals<br />

Chae Un Kim, Buz Barstow, Mark Tate, and Sol Gruner<br />

Research <strong>High</strong>lights<br />

Page 36 CHESS News Magazine 2009


New Mechanical Testing Methods for Structural<br />

Materials at Small Size Scales<br />

Matthew P. Miller<br />

Sibley School of Mechanical and Aerospace Engineering, <strong>Cornell</strong> <strong>University</strong><br />

Abstract<br />

This paper describes new experiments<br />

designed to characterize the<br />

micromechanical response of structural<br />

materials. As a way to calibrate<br />

microscale constitutive models and to<br />

understand grain scale elastic-plastic<br />

deformation, our group worked closely<br />

with CHESS personnel to develop a high<br />

energy x-ray method for measuring<br />

lattice Strain Pole Figures (SPFs) in<br />

situ in the A2 experimental station at<br />

CHESS. Results from experiments and<br />

crystal-based finite element simulations<br />

on copper are presented in the paper.<br />

The distribution of stresses predicted<br />

by the model closely compares to<br />

the experimental result. Motivated<br />

by our success at CHESS, we have<br />

begun conducting experiments at<br />

the Advanced Photon <strong>Source</strong> (APS) to<br />

characterize individual grains within a<br />

deforming polycrystal. We show results<br />

for crystals in two processed states<br />

of a titanium alloy,Ti-7Al. The stress<br />

states change with deformation-likely<br />

due to translation of the stress state<br />

along the single crystal yield surface. A<br />

complicated residual stress state was<br />

present in each grain upon unloading.<br />

By highly resolving individual diffraction<br />

spots within each crystal, we were<br />

able to understand the character of<br />

dislocation structure formation in<br />

each alloy state. These experiments<br />

show huge potential as a means of<br />

characterizing engineering materials in<br />

general. We also believe the experiments<br />

– coupled with a crystal-based stress<br />

state representation methodology –<br />

bring a new measurement capability to<br />

the important problem of residual stress.<br />

Introduction<br />

Structural materials play a key role in<br />

the products we use every day. From the<br />

titanium of a bicycle frame to the nickelbased<br />

super alloy that makes up a jet<br />

engine component, structural materials<br />

are selected for their strength, stiffness,<br />

density, corrosion resistance and a<br />

myriad of other properties. Mechanical<br />

properties are always related to some<br />

idealization of material behavior referred<br />

to as a constitutive model. From Young’s<br />

modulus to yield strength and fracture<br />

toughness, mechanical properties<br />

arise naturally when the response of<br />

a structural material to thermal and<br />

mechanical loading is viewed from the<br />

perspective of a constitutive model.<br />

Mechanical properties are measured<br />

by applying thermo-mechanical loads<br />

to a test specimen then monitoring<br />

its response. By knowing both the<br />

stimulus and response functions, the<br />

properties can be determined. With<br />

the mechanical properties known,<br />

these experiments – which are referred<br />

to as mechanical tests – can also be<br />

used to monitor the performance<br />

characteristics of a structural material.<br />

The fidelity of the constitutive model<br />

can be verified by subjecting the<br />

material to more complicated loading<br />

scenarios. The simplest test is uniaxial<br />

monotonic loading – pull or compress<br />

the specimen past the yield strength to<br />

failure. Multiaxial loading experiments,<br />

such as tension-torsion tests, can be<br />

used to understand material behavior<br />

and the validity of the model under<br />

complicated stress states, which more<br />

closely mimic service conditions. <strong>High</strong><br />

temperature, corrosive conditions,<br />

cyclic loading and loading rate variation<br />

can be employed to create a veritable<br />

“gauntlet” of service conditions for the<br />

material and the model. Success in<br />

correlating experimental behavior builds<br />

confidence that a constitutive model<br />

can be trusted in the actual conditions<br />

an engineering component will be<br />

subjected to – conditions that cannot be<br />

replicated in the laboratory.<br />

Constitutive model development for<br />

structural applications has moved<br />

down scale for several reasons. Microelectromechanical<br />

systems, MEMS, are<br />

the same size as the microstructure.<br />

MEMS design requires an understanding<br />

of material response on the scale of the<br />

microstructure. More importantly, there<br />

is a growing need to model deformation<br />

– induced microstructure evolution<br />

within large scale components. The hope<br />

for these microscale formulations– from<br />

models of atomic behavior to dislocation<br />

models to crystal plasticity – is that some<br />

of the behaviors that are mysterious and<br />

practically random on the macroscale<br />

will become tractable and predictable<br />

when the microstructural response<br />

is being modeled. These processes<br />

include fatigue microcrack initiation,<br />

plasticity, recrystallization and phase<br />

transformations. Enormous growth<br />

has taken place in the general field of<br />

multiscale material modeling over the<br />

past 10-15 years. The ever-increasing<br />

capacity of computing facilities has<br />

encouraged theoreticians to model<br />

behaviors on increasingly smaller size<br />

scales with the hope of creating truly<br />

predictive links between material<br />

structure, thermo-mechanical properties<br />

and measures of performance. These<br />

links will eventually enable designers<br />

to base allowable (operating loads<br />

and temperatures, for instance) on<br />

micromechanical properties and<br />

analyses. From an experimental<br />

perspective, the question becomes<br />

whether or not we can reproduce<br />

mechanical tests on the microscale.<br />

One approach is to conduct traditional<br />

mechanical tests on subsized specimens<br />

constructed from the microstructure or<br />

from components of a micromechanical<br />

machine. Another approach is to<br />

employ microscopic mechanical<br />

instruments such as nanoindentation<br />

devices to probe microscale stress-strain<br />

behavior. <strong>High</strong> energy synchrotron<br />

x-ray diffraction methods represent<br />

a formidable alternative to subscale<br />

experiments. Modern area detectors and<br />

impressive depths of penetration have<br />

CHESS News Magazine 2009 Page 37


made it possible to probe a metallic<br />

crystal that lies far beneath the surface<br />

of a test specimen and make diffraction<br />

measurements quickly. By conducting<br />

carefully designed mechanical tests<br />

in situ during an x-ray diffraction<br />

experiment, the crystal lattice plays<br />

the role of the engineering strain gage<br />

on the macroscale, with the added<br />

bonus that the strains indicated by the<br />

lattice are elastic. The challenge is the<br />

deconvolution of the diffraction data -<br />

pulling it apart in order to understand<br />

its origin and significance – then<br />

recombining it to draw conclusions<br />

regarding the deformation of the<br />

polycrystalline aggregate.<br />

Our group at <strong>Cornell</strong> began using high<br />

energy x-rays at the A2 experimental<br />

station at CHESS to understand<br />

the behavior of polycrystalline<br />

structural materials and to validate<br />

micromechanical constitutive models.<br />

Working closely with beamline<br />

personnel, we designed and fabricated<br />

our own mechanical loading stage<br />

and have had significant success<br />

interrogating polycrystalline aggregates<br />

for distributions of lattice strain and<br />

stress 1,2 . Since that time, we have<br />

moved farther downscale to probe<br />

individual crystals within an aggregate<br />

at the Advanced Photon <strong>Source</strong> (APS).<br />

This paper describes our experiment<br />

developed at A2 designed to measure<br />

lattice strain distributions within a<br />

loaded polycrystalline aggregate. By<br />

determining the lattice strain tensor<br />

at each orientation, we are able to use<br />

Hooke’s law to calculate the orientationaveraged<br />

stress tensor distribution.<br />

Coupled with crystal-based simulations,<br />

these data have proven invaluable<br />

for improving our understanding<br />

of deformation partitioning in<br />

polycrystalline alloys. The second type<br />

of experiment, in which the deformation<br />

of individual crystals within the<br />

aggregate is tracked, is part of the <strong>High</strong><br />

<strong>Energy</strong> Diffraction Microscopy (HEDM)<br />

suite of experiments at beamline 1-IDC<br />

at the APS. When full spatial resolution<br />

is attained, the HEDM experiment will<br />

likely become the “gold standard” for<br />

microscale stress measurements.<br />

Stress State Distributions in<br />

Loaded Polycrystals<br />

In this "powder" experiment , we<br />

interrogate an entire polycrystalline<br />

aggregate simultaneously. As with<br />

traditional mechanical tests, each<br />

continuum point within the sample<br />

is identical. Each diffraction volume<br />

that defines a continuum point must<br />

contain a statistically relevant number of<br />

grains - usually several thousand. In this<br />

way, we are not restricted to tracking<br />

a particular set of crystals during the<br />

in situ deformation. This condition is<br />

crucial for this experiment and dictates<br />

the grain size / beam size relationship.<br />

We acquire bands of diffraction data for<br />

each family of lattice planes, {hkl}, by<br />

rotating the loadframe using its builtin<br />

goniometer 3 . At A2 we typically use<br />

a beam of 50 KeV x-rays with 0.5mm x<br />

0.5mm slits. A set of lattice Strain Pole<br />

Figures (SPFs) are measured at discrete<br />

loads; then the lattice strain distribution<br />

tensor, є(R), at each lattice orientation,<br />

R, is determined 4 . Lattice strains are<br />

elastic so the stress distribution is<br />

calculated using Hooke’s law, σ(R) =<br />

C(R)є(R), where C are the single crystal<br />

moduli. We know that orientation as<br />

well as crystallographic neighborhood<br />

contribute to the mechanical response of<br />

an individual crystal within an aggregate.<br />

Therefore σ(R) has the interpretation of<br />

being the most likely state of stress at a<br />

particular orientation. The σ(R) data are<br />

ideal for use with a crystal-based modeling<br />

framework, such as an elastic-plastic finite<br />

element formulation, for understanding<br />

the distribution of deformation within<br />

the aggregate. One way to evaluate the<br />

distribution of each stress component over<br />

orientation space - both measured and<br />

computed - is by an expansion of spherical<br />

∞<br />

harmonic functions, σij(R) = Ʃ H l T l (R).<br />

l=0<br />

Here the T l (R) are generalized spherical<br />

harmonics functions (modes) that satisfy<br />

relevant symmetry conditions and form<br />

an orthonormal basis over orientation<br />

space. The coefficients, H l , prescribe the<br />

contribution of each mode to the stress.<br />

Fig. 1 depicts two spherical harmonic<br />

modes plotted over orientation space.<br />

Fig. 1 also depicts the spherical harmonics<br />

coefficients for the shear stress, σ xy , over<br />

orientation space from a polycrystalline<br />

copper sample loaded in tension to a<br />

macroscopic stress of S zz = 180 MPa. Based<br />

upon the coefficients, we see that both the<br />

experiment and simulation have “chosen”<br />

mode 7 as the dominant mode. Even<br />

though this component of stress must<br />

necessarily integrate to zero over the cross<br />

section to match the boundary conditions<br />

of the experiment, shear stress values as<br />

large as 75-80 MPa are present. The close<br />

comparison between the simulation and<br />

experiment creates confidence in both,<br />

and builds additional trust in the model<br />

as it is applied beyond the reach of the<br />

experiments. Similar comparison was made<br />

(a) mode 3 (b) mode7 (c) σ xy<br />

Fig. 1: (Left and Center)Two spherical harmonic modes shown distributed over the cubic fundament region of orientation space parameterized using the<br />

Rodrigues representation of orientation. The surface of the region along with orthogonal slices through the origin are depicted. (Right) Spherical Harmonics<br />

(SH) coefficients vs. mode number for the shear stress component, σ xy as determined from the experimental data (EXP) and computed from the crystalbased<br />

finite element simulations (SIM) at a uniaxial stress of S zz = 180 MPa. Three finite element mesh sizes for the virtual test specimen were employed,<br />

1,098, 2,916 and 10,976 crystals. Each crystal contained 48 finite elements.<br />

Page 38 CHESS News Magazine 2009


for other components of stress 5 . From<br />

these experiments and simulations,<br />

we learned that the microscale stress<br />

state can be very different than the<br />

macroscopically applied stress. There<br />

are many situations (such as fatigue)<br />

when an assumption of uniaxial stress<br />

is nonconservative. These data enable<br />

us to implement the finite element<br />

modeling formulation with greater<br />

confidence.<br />

Individual Crystal Experiments<br />

The stress distribution data from the<br />

in situ SPF experiments described<br />

in the previous section provide an<br />

orientation dependent, statistically<br />

relevant approximation of the<br />

micromechanical partitioning of the<br />

elastic-plastic deformation over a<br />

polycrystalline aggregate. As described<br />

above, the stress distribution data<br />

are an excellent match for the current<br />

generation of crystal based simulation<br />

formulations. In reality, the response<br />

of each crystal within the aggregate<br />

is a function of both its orientation<br />

and the surrounding crystals that<br />

supply the boundary conditions for<br />

its deformation. Measurement of the<br />

stress state within individual crystals<br />

will enable us to eventually quantify the<br />

neighborhood effect. The 3DXRD suite<br />

of interrogation experiments were first<br />

developed by Risoe researchers at the<br />

European <strong>Synchrotron</strong> Radiation Facility<br />

(ESRF) 6 . One of the 3DXRD capabilities<br />

made it possible to measure lattice<br />

strains within an individual crystal<br />

within a loaded aggregate 7 . Our group<br />

has collaborated with APS sector 1<br />

personnel to further develop this in situ<br />

lattice strain measurement capability<br />

as a part of the beamline 1-IDC suite<br />

of experiments called <strong>High</strong> <strong>Energy</strong><br />

Diffraction Microscopy (HEDM). Fig. 2<br />

depicts a configuration for conducting<br />

two HEDM experiments: individual<br />

crystal lattice strain measurements<br />

for understanding crystal stress states<br />

and high resolution interrogation of<br />

individual reflections within the crystal<br />

to understand deformation-induced<br />

dislocation substructure formation 8 .<br />

Titanium Alloys<br />

With their high strength to density<br />

ratio and excellent corrosion resistance<br />

and temperature range, titanium<br />

alloys have enabled some of the most<br />

important aircraft design innovations<br />

over the past several decades 9 . Single crystal property anisotropies make processing of<br />

titanium alloys difficult and create challenging design conditions, however. Our group<br />

has conducted a number of studies on titanium. Fig. 2 also depicts the macroscopic<br />

stress strain curves for experiments on an α phase (hcp) Ti-7Al alloy in two states. The<br />

slow cooling rate that the AC (Air Cooled) material experiences from α / ß transus<br />

temperature produces an ordering of the titanium and aluminum atoms over short<br />

length scales. Dislocation glide within this Short Range Order (SRO) material results in<br />

planar slip and increased strain hardening and greatly affects the room temperature<br />

creep behavior of Ti-7Al 10 . When the Ti-7 was quenched quickly in the IWQ (Ice Water<br />

Quenched) state, the SRO does not form and the slip is more isotropic. We conducted<br />

HEDM experiments - lattice strain measurements and high resolution reflection<br />

decomposition - on samples from both material states. The stress strain curves shown<br />

in Fig. 2 also depict the macroscopic stress levels where diffraction experiments were<br />

conducted. We tracked the deformation of several grains in each specimen but only<br />

collected high resolution data on one each. While identical orientations would have<br />

been preferred, we chose the subject grains based mostly on the number of clear<br />

reflections we were able to capture. The angles between the c axis and the loading<br />

axis in the IWQ and AC grains was 28.7°and 81.0°, respectively. Monochromatic 52 KeV<br />

x-rays were employed with a beam size of 300 μm x 150 μm (horizontal x vertical) for<br />

detector A and 100 μm x 100 μm for detector B. Additional experimental details can be<br />

found in 11 .<br />

Fig. 2: (Left) HEDM setup at APS beamline 1-IDC. Different detectors were used for the lattice strain<br />

measurements and for the high resolution experiments. Each experiment requires a rotation of Φ about<br />

the y axis to interrogate relevant scattering vectors. Detector A (approximately 1m from the sample)<br />

captures many discrete diffraction spots from multiple {hkl}s and the lattice strains are used to determine<br />

the full lattice strain tensor for the grain. A single reflection is captured on detector B - very high resolution<br />

is possible. The B detector is approximately 8m from the sample. The x', y', z' laboratory and q x', q y', q z'<br />

reciprocal space coordinate systems are indicated. The y' axis coincides with the sample y direction,<br />

the sample x and z directions rotate with Φ. The insert indicates diffraction from a selected bulk grain.<br />

(Right) Macroscopic stress-strain response for the AC and IWQ Ti-7Al samples during the in situ loading<br />

experiment. The symbols indicate stress levels where diffraction experiments were conducted.<br />

Stress State Evolution<br />

Fig. 3 depicts the evolution of the three dimensional stress state within the AC and<br />

IWQ grains. Each grain begins with a stress state that very nearly matches the applied<br />

uniaxial state. As the load increases and the grains yield, the stress states change.<br />

The largest change in each grain’s stress state occurs between points 2 and 3, when<br />

fully developed plasticity occurs. The stress states become multiaxial - which is very<br />

interesting from an elastic-plastic deformation perspective and validates, at least<br />

in a qualitative manner, the trends that we observed in the aggregate experiments<br />

described above. The direction associated with the largest principal stress is near<br />

the macroscopic loading direction, but the other two principal components are not<br />

zero, which they would be for a uniaxial stress state. Each specimen was unloaded to<br />

400 MPa, which corresponds to the first stress level. The residual stress state in each<br />

grain can be understood by comparing 400el values to 400un values. Even “simple”<br />

macroscopic loading conditions like uniaxial tension can produce complex, threedimensional<br />

residual stress states on the crystal scale that evolve with straining. Since<br />

most residual stress measurement capabilities require an assumption of stress state 12 ,<br />

this indicates a real need for a more sophisticated measurement capability. Processinginduced<br />

residual stresses are a huge issue in titanium parts used in aerospace<br />

CHESS News Magazine 2009 Page 39


applications. For instance, it is nearly impossible to predict the fatigue life of a titanium part that contains significant residual<br />

stresses.<br />

Resolved Shear Stress<br />

Dislocation glide (yielding on the crystal scale) occurs when the shear stress on one or more of the slip systems (slip plane + slip<br />

direction) reaches a critical value. One of the primary hcp slip systems is the basal, {0001} 〈1120〉.<br />

−<br />

Calculation of the resolved shear<br />

stress, τ , involves projecting the stress state onto the slip plane in the slip direction. The typical approach is to assume uniaxial<br />

tension as the macroscopic stress state. We have measured the full stress tensor so we do not need to assume a uniaxial stress<br />

state. To understand the possible error that might arise with a uniaxial assumption, we calculated τ in the AC grain two ways. First<br />

we employ the usual approach: project the macroscopic stress, S yy , onto the slip system. Then we calculate τ using the entire stress<br />

tensor. The results are depicted in Fig. 3. The biggest difference is seen between 595 MPa and 630 MPa where fully developed<br />

plasticity ensues and the stress state jacks in Fig. 3 are seen to rotate. We actually see a drop in τ between 595 MPa and 630 MPa for<br />

two of the three slip systems, even though the macroscopic stress increases by 35 MPa. This seemingly inconsistent result arises<br />

due to the topology of the single crystal yield surface. Metallic crystals are restricted to plastic straining only on the prescribed slip<br />

systems. To accommodate a general plastic strain rate, therefore, it has been hypothesized that the crystal stress state will actually<br />

rotate to activate more slip systems 13,14 . Certainly this evolution can involve a decrease in the value of stress projected onto a slip<br />

system. The decrease seen in Fig. 3 is entirely consistent with this rotation of the stress state. To our knowledge, this is the first<br />

measurement of the yield surface-induced stress rotation for a crystal embedded within a polycrystalline aggregate. The rotation<br />

of the stress state should not be confused with the crystal rotation, which is discussed in the next section.<br />

Fig. 3: (Left)Principal axis triads or<br />

”jacks” depicting the orientation<br />

of the principal stress state at the<br />

four indicated macroscopic stress<br />

levels (in MPa) for the IWQ (top)<br />

and AC (bottom) specimen. The<br />

legs of the jacks correspond to<br />

the principal directions relative to<br />

the specimen Loading Direction<br />

(LD), Transverse Direction (TD)<br />

and Normal Direction (ND). The<br />

colors of each leg and accompanying color scale indicate the principal value. (Right) Resolved shear stress, τ for the 3 basal slip systems of the AC grain as a<br />

function of the applied load. The box symbols were calculated assuming a uniaxial stress and the macroscopic stress value. The X symbols were calculated<br />

using the full stress tensors shown in the figure on the left.<br />

<strong>High</strong> Resolution Reflection Decomposition<br />

We hypothesized that the SRO-induced differences between AC and IWQ would be visible in the high resolution data. Using<br />

detector B, we analyzed individual reflections (spots) associated with the AC and IWQ grains. Basically this study is an extension of<br />

traditional peak broadening analysis. The detailed geometric information that can be obtained with high brilliance synchrotron<br />

radiation and large sample to detector distances however, enable increased understanding of the deformation-induced<br />

dislocation substructure formation that occurs within metallic crystals.<br />

Fig. 4 depicts reciprocal space representations of intensity and lattice strain distribution for individual reflections in the AC<br />

and IWQ grains. Fig. 4(a) are intensity maps. An interesting feature of the AC map is the extended “tail” which is consistent with<br />

inhomogeneous grain rotation. The intensities integrated over the peak and tail region are roughly the same; hence, the volumes<br />

are similar 15 . This would indicate significant dislocation structure formation, which is consistent with TEM micrographs of the<br />

material 11 . The intensity distribution for the IWQ reflection is consistent with relatively small grain rotation and homogeneous<br />

deformation and distribution of dislocations. Fig. 4(b) depicts the “centroid” of the axial distributions, q y' . The centroid values<br />

represent the lattice strain in the direction of q y' averaged over the orientation fiber specified by the q x' and q z' coordinates. There<br />

is a broader spread in centroid over the AC grain than the IWQ grain. Fig. 4(c) depicts intensity vs. q y' for the maximal values in the<br />

centroid maps and the peak intensity point from the azimuthal map. The multiple peaks (black, red and blue) in the AC data are<br />

consistent with heterogeneous straining within the grain. From the profiles, the IWQ grain appears to be deforming uniformly.<br />

Additional data from multiple diffraction peaks would be necessary to quantify the exact distribution of orientations within the<br />

grain - an intragrain Orientation Distribution Function, but the information from the reflections shown here is consistent with<br />

the hypothesis regarding short range order and planar slip. From the TEM analysis, we determined that the dislocation density<br />

in the AC grain was comparable to that in the IWQ 11 . We used the IWQ integrated profile (green curve) and the restricted random<br />

dislocation model developed by Wilkins 16 to confirm the dislocation density. However, an estimate of the AC grain dislocation<br />

density from its integrated peak information would grossly overestimate the dislocation density in the AC grain. The broadening<br />

in this case comes primarily from the larger structures associated with heterogeneous straining within the grain.<br />

Page 40 CHESS News Magazine 2009


Fig. 4: Reciprocal space maps<br />

and intensity profiles for<br />

individual reflections from the<br />

AC and IWQ crystal. (a) Azimuthal<br />

map of diffracted intensity in<br />

reciprocal space. The reciprocal<br />

space coordinates are depicted in<br />

Fig. 2. The dashed lines and Miller<br />

indices indicate the directions of<br />

the smallest and largest widths<br />

at half maximum. (b) Azimuthal<br />

map of the centroid of the q y'<br />

(d −1 ) distribution. (c) Normalized<br />

intensity vs. q y' for points 1 and 2<br />

in (b) and for the peak intensity<br />

value in (a).<br />

A: AC (1212)<br />

− − B: IWQ (0220)<br />

−<br />

Summary and Future Directions<br />

Motivated by traditional mechanical<br />

tests, we have conducted experiments<br />

that combine mechanical loading<br />

with synchrotron x-ray diffraction<br />

to understand the micromechanical<br />

response of metallic alloys.<br />

• The aggregate-based experiments<br />

that we developed at CHESS A2 have<br />

become our standard mechanical<br />

test. We have characterized copper,<br />

titanium, aluminum and nickel<br />

alloys. In addition to building a more<br />

complete understanding of our error<br />

bars and resolution, we are continuing<br />

to develop data reduction schemes<br />

that have greatly streamlined the SPF<br />

experiments and made them more<br />

user friendly. In general, any success we<br />

have enjoyed with these experiments<br />

is due in large part to the environment<br />

of encouragement that exists at CHESS<br />

and the excellent working relationship<br />

we have with CHESS personnel. Our<br />

graduation to other techniques and<br />

facilities has come about by the<br />

confidence we have gained from our<br />

CHESS experience.<br />

• In addition to understanding elasticplastic<br />

deformation behavior within a<br />

polycrystalline aggregate, we foresee the<br />

combined suite of distribution-based<br />

SPF diffraction experiments and crystalbased<br />

micromechanical representations<br />

as a way to understand residual stresses in<br />

a way that has never been done. Residual<br />

stress plays a major role in the processing<br />

and performance of wrought metallic<br />

alloys. As demonstrated in this paper,<br />

the state of stress within an aggregate is<br />

a function of macroscopic loading and<br />

microscopic mechanical properties. In<br />

order to accurately quantify and predict<br />

a residual stress state, one must have<br />

a way for accounting for both. Our<br />

experiments and simulations represent a<br />

potential framework capable of such an<br />

accounting.<br />

• In addition to monotonic loading,<br />

we conduct cyclic experiments -<br />

determining SPFs after a certain number<br />

of cycles. In collaboration with Peter<br />

Revesz at CHESS, we have also developed<br />

an experiment for measuring lattice<br />

strains in real time using a custom built<br />

x-ray shutter system 17 .<br />

• The potential of the HEDM experiments<br />

is enormous. We are currently working<br />

with other 1-IDC users to transform the<br />

experimental suite from “heroic efforts”,<br />

capable of measuring the response of<br />

a handful of grains, to methodologies,<br />

capable of characterizing large<br />

polycrystalline aggregates of grains.<br />

Creating robust data acquisition and<br />

reduction software is at the heart of<br />

this effort. Plans include combining the<br />

lattice strain measurements described<br />

above with grain reconstruction<br />

experiments 18 to characterize spatially<br />

resolved stress fields over a statistically<br />

relevant number of grains. Again, the<br />

exciting part of this plan is combining<br />

such datasets with micromechanical<br />

modeling formulations.<br />

Acknowledgments<br />

The results presented here form parts<br />

of the Ph.D. dissertation research of Dr.<br />

Joel V. Bernier and Dr. Jun-Sang Park.<br />

Other members of the group include<br />

Ph.D. students Jay C. Schuren and Kevin<br />

P. McNelis and post-doctoral associate Dr.<br />

Christos Efstathiou. The simulations were<br />

conducted by Professors Paul Dawson of<br />

CHESS News Magazine 2009 Page 41


<strong>Cornell</strong> and Tong Han of Yonsei, Korea. The collaborations with beamline scientists Dr. Alexander Kazimirov (CHESS) and Dr. Ulrich<br />

Lienert (APS) were crucial to this research. Our external collaborators on the Ti-7Al experiments were Professor Michael Mills of<br />

The Ohio State <strong>University</strong> and Dr. Matthew Brandes of the Naval Research Laboratory. The research has been supported financially<br />

by the Air Force Office of Scientific Research, #F49620-02-1-0047, the National Science Foundation, grant #CMS0301615 and the<br />

Office of Naval Research, grant # N00014-05-1-0505.<br />

References<br />

1. M.P. Miller, J.V. Bernier, J.-S. Park, and A. Kazimirov; "Lattice Strain Pole Figure Measurements using <strong>Synchrotron</strong> X-rays", CHESS<br />

News Magazine, pages 44–47 (2005)<br />

2. M.P. Miller, J.V. Bernier, J.-S. Park, and A. Kazimirov; "Experimental Measurement of Lattice Strain Pole Figures using <strong>Synchrotron</strong><br />

X-rays", Review of Scientific Instruments 76:113903 (2005)<br />

3. J.C. Schuren, S. Watts, J.-S. Park, and M.P. Miller; "A System for Measuring Crystal Level Stresses in Deforming Polycrystals", In<br />

Proceedings of the 2007 SEM Annual Conference and Exposition on Experimental and Applied Mechanics, sec. 82 p4, Bethel,<br />

Connecticut, Society for Experimental Mechanics, Inc. (2007)<br />

4. J.V. Bernier and M.P. Miller; "A Direct Method for the Determination of the Mean Orientation Dependent Elastic Strains and Stresses<br />

in Polycrystalline Alloys from Strain Pole Figures", Journal of Applied Crystallography, 39:358–368 (2006)<br />

5. M.P. Miller, J.-S. Park, P.R. Dawson, and T.-S. Han; "Measuring and Modeling Distributions of Stress State in Deforming Polycrystals",<br />

Acta Materialia, 56:3827–3939 (2008)<br />

6. H. Poulsen; "Three-Dimensional X-Ray Diffraction Microscopy", Springer, Heidelberg, U.K. (2004)<br />

7. L. Margulies, H.F. Poulsen T. Lorentzen, and T. Leffers; "Strain Tensor Development in a Single Grain in the Bulk of a Polycrystal<br />

Under Loading", Acta Materialia, 50(7):1771–1779 (2002)<br />

8. B. Jacobsen, H.F. Poulsen, U. Lienert, J. Almer, S.D. Shastri., H.O. Sorensen, C Gundlach, and W. Pantleon; "Formation and<br />

Subdivision of Deformation Structures During Plastic Deformation", Science, 312:889–912 (2006)<br />

9. G. Lutjering and J. Williams; "Titanium (Engineering Materials and Processes)", Springer Verlag, Berlin (2003)<br />

10. T. Neeraj and M.J. Mills; "Short-Range Order (sro) and Its Effect on the Primary Creep Behavior of a ti – 6wt", Materials Science and<br />

Engineering A, A319-321:415–419 (2001)<br />

11. U. Lienert, M. Brandes, J.V. Bernier, J. Weiss, S. Shastri, M.J. Mills, and M.P. Miller; "In-situ Single Grain Peak Profile Measurements<br />

on ti-7al During Tensile Deformation", Materials Science A, In press (2009)<br />

12. Society of Experimental Mechanics. Handbook of Measurement of Residual Stresses. Fairmont Press, Lilburn, Georgia, USA<br />

(1996)<br />

13. U.F. Kocks; "The Relation Between Polycrystal Deformation and Single Crystal Deformation", Metallurgical Transactions,<br />

1:1121–1143 (1970)<br />

14. H.R. Ritz, P.R. Dawson, and T. Marin; "Analyzing the Orientation Dependence of Stresses in Polycrystals using Vertices of the Single<br />

Crystal Yield Surface", Journal of the Mechanics and Physics of Solids, page In Review (2009)<br />

15. B.D. Cullity; "Elements of X-Ray Diffraction", Addison-Wesley, Reading, MA, 2nd edition (1978)<br />

16. M. Wilkins; "The Determination of Density and Distribution of Dislocations in Deformed Single Crystals from Broadened X-ray<br />

Diffraction", Physical Status Solidi (A), 2:359–370 (1970)<br />

17. J.-S. Park, P. Revesz, A. Kazimirov, and M.P. Miller; "A Methodology for Measuring in situ Lattice Strain of Bulk Polycrystalline<br />

Material under Cyclic Load", Review of Scientific Instruments, 78:023910 (2007)<br />

18. R.M. Suter, C.M. Heffernan, S.F. Li, D. Hennessy, and C. Xiao; "Probing Microstructure Dynamics with X-ray Diffraction<br />

Microscopy", Journal of Engineering Materials and Technology, 130:021007 (2008)<br />

Page 42 CHESS News Magazine 2009


Alteration of Citrine Structure by Hydrostatic Pressure<br />

Sol M. Gruner 1,2,3 , Buz Barstow 1 , Nozomi Ando 2 , and Chae Un Kim 2,3<br />

1<br />

School of Applied Physics, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

Department of Physics, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

In 1914 Percy Bridgman, the Nobel<br />

prize winning father of high pressure<br />

physics, squeezed hen egg white at high<br />

pressure and found that it coagulated,<br />

as though it had been hard-boiled 1 . He<br />

had discovered that proteins denature<br />

under pressure, resulting in a coagulated<br />

state reminiscent of what happens when<br />

proteins are thermally denatured. Since<br />

Bridgman’s pioneering observation<br />

many hundreds of papers have reported<br />

on the effects of high pressures on<br />

biomolecular systems: Proteins unfold,<br />

enzymatic rates change, multimeric<br />

assemblies have greatly altered<br />

stabilities, viral infectivities are greatly<br />

different, etc. Frequently, the changes<br />

are of very large magnitude and occur at<br />

pressures encountered in the biosphere.<br />

Yet proteins are very incompressible,<br />

typically 10 times stiffer than water,<br />

so the actual change in volume is<br />

very small, even at a few thousand<br />

atmospheres pressure. What are the<br />

mechanisms of the pressure effects<br />

What do these mechanisms teach us<br />

about protein function<br />

Little is known about structural changes<br />

in proteins in response to pressure and<br />

the way the resultant changes affect<br />

function. The main reason for this<br />

lack of understanding is that very few<br />

protein structures have been solved as<br />

a function of pressure. Another reason<br />

is a cultural bias that pressure effects<br />

are esoteric and have little relevance to<br />

understanding proteins. This thinking is<br />

misguided for two reasons: First, there<br />

are significant pressure effects observed<br />

in the biosphere. For example, certain<br />

essential enzymatic processes adapted<br />

for deep sea life fail at atmospheric<br />

pressure, and vice versa. Second,<br />

pressure is another thermodynamic<br />

“knob” that can be turned by the<br />

experimenter to gain insight about<br />

protein function. Put another way, the<br />

free energy that governs molecular<br />

reactions is related to d(PV – TS), where P<br />

is pressure, V is volume, T is temperature<br />

and S is the entropy. In many protein systems, a change in pressure of a thousand<br />

atmospheres is comparable in magnitude to a change of many tens of degrees in<br />

temperature.<br />

In recent years we have developed methods of data acquisition and analysis<br />

to perform protein crystallography at pressures ranging to several thousand<br />

atmospheres in order to study pressure-induced structural changes 2,3,4 . The studies<br />

have provided a wealth of information on phenomena, including pressure-induced<br />

protein unfolding 5 , water penetration into proteins 4,6 , and conformational changes<br />

that affect function 2,7 . A general theme that emerges is that pressure induces a host<br />

of structural displacements at the level of a few tenths of an angstrom. Although<br />

these displacements are well below typical resolution limits for observation of single<br />

atoms in proteins, they are readily observed for collections of protein residues and<br />

secondary structures by observing changes in the center of mass of the protein parts.<br />

This should not be surprising – recall that changes in displacements of atomic force<br />

microscope cantilevers are routinely observed at the fraction of an angstrom level,<br />

even though the resolution of the visible light optical systems involved is only a few<br />

tenths of a micron.<br />

Why are these small changes significant Most enzymatic reactions involve<br />

conformational changes of only a few tenths of an angstrom at active sites. Therefore,<br />

changes of this magnitude can distort active sites resulting in big effects on reaction<br />

rates. Observation of the structural changes and consequent effects on reaction<br />

rates provides important insight on the detailed structural requirements of enzyme<br />

reactive sites.<br />

A recent study at CHESS used high pressure crystallography to understand the<br />

pressure sensitivity of fluorescence in the protein Citrine 7 . Citrine is a modified<br />

green fluorescent protein (GFP) whose peak fluorescent wavelength is known to<br />

shift with pressure. All proteins in the GFP family have a β-barrel structure with a<br />

fluorescent chromophore at the center, formed by the autocatalytic fusion of three<br />

amino acid residues (Fig. 1). In Citrine, an additional tyrosine ring is stacked upon<br />

the chromophore. The fluorescence peak of Enhanced GFP (MEGFP) is not pressure<br />

dependent to at least a few thousand atmospheres (Fig. 2) (wild type GFP does<br />

Fig. 1: Ribbon diagram of<br />

Citrine showing the residues<br />

that make up the chromophore<br />

in the core of the protein 7 .<br />

Fig. 2: Change of the peak fluorescent<br />

wavelength with pressure for Citrine (red<br />

squares) and a modified GFP (blue diamonds) 7 .<br />

CHESS News Magazine 2009 Page 43


display pressure sensitive behavior, for<br />

reasons that are related to the reduction<br />

in fluorescence intensity of Citrine rather<br />

than the shift in fluorescence peak).<br />

Basic quantum chemistry suggests that<br />

the peak fluorescent wavelength of a<br />

stacked aromatic system would be very<br />

sensitive to the detailed coupling, e.g.,<br />

the relative positions of the π-bonds<br />

of the aromatic rings. Since GFP has a<br />

single aromatic chromophore, it would<br />

be relatively insensitive to movements<br />

of the chromophore with pressure. On<br />

the other hand, Citrine, with a stacked<br />

aromatic chromophore, would be very<br />

sensitive to small displacements of one<br />

aromatic ring relative to the other. Our<br />

study sought to observe if there is a<br />

relative displacement of the expected<br />

magnitude with pressure. If so, this may<br />

be taken as generally analogous for<br />

what happens in active sites in enzymes.<br />

That is to say, comparable distortions<br />

of active sites can be expected to alter<br />

reaction rates, thereby explaining in a<br />

general way why enzymes are pressure<br />

sensitive.<br />

The relative motions versus pressure<br />

of the two aromatic rings are shown<br />

in Fig. 3. What gives rise to these<br />

displacements The two aromatic rings<br />

that make up Citrine’s chromophore<br />

are connected to very different parts<br />

of the single polypeptide chain from<br />

which Citrine is folded. One of the rings<br />

hangs off the inside of the β-barrel<br />

while the other is in another part of the<br />

polypeptide chain that threads through<br />

the axis of the barrel. Thus, distortions<br />

of the overall structure are sensitively<br />

coupled to relative displacements<br />

between the stacked rings. Specifically,<br />

the residues that make up Citrine can<br />

be grouped into two clusters. Cluster<br />

1 makes up one side of the barrel<br />

and Cluster 2 includes the threading<br />

polypeptide chain. A careful analysis of<br />

the changes in overall Citrine structure<br />

with pressure shows that Cluster 1<br />

contracts, resulting in a bending motion<br />

relative to residues of Cluster 2. The<br />

relative motions of the two clusters<br />

result in relative motions of the two<br />

rings.<br />

This study illustrates how collective<br />

distortions of the overall structure of a<br />

protein are conveyed to critical parts of<br />

the structure necessary for “function”,<br />

“function” in this case being peak<br />

fluorescent wavelength. Citrine was<br />

chosen for this study because there<br />

is good fundamental understanding<br />

of how distortions of the stacked<br />

aromatic rings should affect the peak<br />

fluorescence. In enzymes, however,<br />

the detailed structural requirements<br />

of active sites are frequently less<br />

well understood. The Citrine study<br />

also suggests how pressure studies<br />

on proteins may guide mutagenic<br />

engineering to optimize enzymatic<br />

function. Given the observed distortions<br />

of the chromophore required to achieve<br />

a given peak fluorescence shift, can<br />

these be used to guide programmed<br />

mutations to generate the shift at<br />

room pressure The study suggests<br />

that insertion of smaller residues in<br />

the barrel side of Cluster 1 might<br />

result in the same kind of contraction<br />

needed to mimic the shift seen at<br />

high pressure. This experiment has yet<br />

to be performed, but it suggests an<br />

overall strategy to optimize enzymes 7 :<br />

As noted earlier, many enzymes have<br />

pressure-dependent catalytic rates. The<br />

strategy would involve observation of<br />

the structural changes that result in<br />

increased activity with pressure. Once<br />

these are identified, one would attempt<br />

Fig. 3: Movement of one part of the chromophore relative to the other with pressure 7 .<br />

to perform targeted mutagenesis to<br />

mimic the structural deformations.<br />

The Citrine study has shown that<br />

the pressure-dependence of protein<br />

function can be understood in<br />

terms of straightforward collective<br />

displacements of protein structure.<br />

Prediction of the exact conformational<br />

changes that occur is beyond<br />

the present state-of-the-art, but<br />

experimental observation of the<br />

changes has been demonstrated.<br />

In the future these may be used to<br />

engineer protein activity.<br />

References:<br />

1. P.W. Bridgman; “The Coagulation<br />

of Albumen by Pressure”, Journal<br />

of Biological Chemistry, 19(4): p.<br />

511-512 (1914)<br />

2. P. Urayama, G.N. Phillips Jr, and S.M.<br />

Gruner; “Probing Substates in Sperm<br />

Whale Myoglobin Using <strong>High</strong>-pressure<br />

Crystallography”, Structure 10: p.<br />

51-60 (2002)<br />

3. C.U. Kim, and S.M. Gruner; “<strong>High</strong><br />

Pressure Cooling of Protein Crystals<br />

without Cryoprotectants”, Acta Cryst.<br />

D61: 881-890 (2005)<br />

4. M.D. Collins, M.L. Quillin, G. Hummer,<br />

B.W. Matthews, and S.M. Gruner;<br />

“Structural Rigidity of a Large Cavitycontaining<br />

Protein Revealed by <strong>High</strong>pressure<br />

Crystallography”, J. Mol. Biol.<br />

367: p. 752-763 (2007)<br />

5. N. Ando, B. Barstow, W.A. Baase,<br />

A. Fields, B.W. Matthews, and<br />

S.M. Gruner; “Structural and<br />

Thermodynamic Characterization<br />

of T4 lysozyme Mutants and The<br />

Contribution of Internal Cavities to<br />

Pressure Denaturation”, Biochemistry<br />

47(42): p. 11097-11109 (2008)<br />

6. M.D. Collins, G. Hummer, M.L.<br />

Quillin, B.W. Matthews, and S.M.<br />

Gruner; “Cooperative Water Filling of<br />

a Nonpolar Protein Cavity Observed<br />

by <strong>High</strong>-pressure Crystallography<br />

and Simulation”, PNAS 102(45): p.<br />

16668-16671 (2005)<br />

7. B. Barstow, N. Ando, C.U. Kim, and<br />

S.M. Gruner; “Alteration of Citrine<br />

Structure by Hydrostatic Pressure<br />

Explains the Accompanying Spectral<br />

Shift”, Proc. Natl. Acad. Sci. USA 105:<br />

p. 13362-13366 (2008)<br />

Page 44 CHESS News Magazine 2009


Putting Color into Surface Diffraction<br />

Detlef-M. Smilgies<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Surface X-Ray Diffraction (SXRD), also called Grazing-Incidence Diffraction (GID) is a well-established powerful tool for structure<br />

determination of surfaces 1,2 , monolayers 3 , and thin films 4,5 . Moreover, surface phase transitions 1,2 and time-dependent effects, such<br />

as growth kinetics 6 , have been studied in depth with this method. However, SXRD/GID also acquired a reputation of being highly<br />

esoteric with very refined representations of scattering data in unusual units.<br />

Since molecular thin films have come into focus, be it for organic electronics device-grade films, self-organized functional<br />

coatings, or biophysical membranes, this spartan picture has begun to change. While in close-to-perfect inorganic structures all<br />

information is contained along the diffuse scattering rods perpendicular to the substrate surface, soft materials inherently have<br />

a larger amount of disorder. Hence, rather than just scanning the rods, now full reciprocal space maps (RSM) have become of<br />

interest, in order to determine structure and distinguish different types of disorder at the same time. Along with the reciprocal<br />

space maps has come the color.<br />

While we originally started out working in the conventional<br />

way of characterizing molecular films by hunting for molecular<br />

Bragg reflections along the symmetry directions of the<br />

substrate 4 , it soon became apparent that this was not the most<br />

efficient route to find all sufficiently strong film reflections, in<br />

order to collect as much structural information as obtainable.<br />

Moreover molecular films are often characterized by lowsymmetry<br />

lattices such as monoclinic and triclinic, which<br />

complicates this search even more.<br />

Hence we started combining the linear detection scheme<br />

from liquid surface diffraction 3 with the oscillation method<br />

from protein crystallography. The oscillation range could<br />

be set either narrow for proper integration of all reflections<br />

along a single scattering rod 7 or wide enough to cover the full<br />

irreducible range of rotation as given by substrate symmetry 8 .<br />

The latter “survey scans” provided the desired efficient method<br />

to identify all strong film reflections. In combination with<br />

azimuth scans around the surface normal at specific q-values<br />

to measure the scattering rods, all available information can be<br />

efficiently collected. We called our method GI-RSM 7 .<br />

GI-RSM proved to be also an effective method for uniaxial<br />

powders. These are organic thin films that are grown on<br />

amorphous substrates such as native or thermal oxides of<br />

silicon wafers, glass, or indium-tin oxide (ITO) coatings. In such<br />

cases there often is a well-defined orientation of the molecular<br />

layers with respect to the surface, however, crystallites are<br />

oriented at random in the surface<br />

plane. The Bragg condition is met<br />

easily, where the ring-like Bragg<br />

reflections oriented parallel to<br />

the surface intercept the Ewald<br />

sphere. Hence 2D powders in<br />

grazing incidence geometry<br />

produce discrete spots in the<br />

scattering images, and the<br />

overlap problem of conventional<br />

powder diffraction is much<br />

reduced. The resulting diffraction<br />

patterns closely resemble fiber<br />

diffraction patterns.<br />

Fig. 1: Radial scan with<br />

oscillation along the [110]<br />

high-symmetry plane of<br />

the KCl substrate (top) and<br />

azimuth scan revealing the<br />

fine structure of the (11L) rod<br />

(bottom). The boomerangshaped<br />

BP1T molecule<br />

containing both phenyl and<br />

thiophene rings and related<br />

molecules can be used<br />

in organic light-emitting<br />

devices covering the color<br />

range from red to blue 7 .<br />

Fig. 2: First incarnation of<br />

the GI-RSM set-up at G2, still<br />

using the horizontal four-circle<br />

diffractometer 7 . After upgrading<br />

the instrument to a six-axis<br />

psi-circle diffractometer 9 the<br />

detector arm can now also<br />

scan in the vertical direction<br />

extending the access to<br />

higher exit angles β, which<br />

was previously limited by the<br />

aperture of the linear detector.<br />

CHESS News Magazine 2009 Page 45


A first success for the Resel group was collecting<br />

enough information from a pentacene film to<br />

compare the structure of the thin film phase<br />

predicted by first principles methods with<br />

these low-resolution diffraction data 10 . By<br />

now we have refined data collection further<br />

with the new kappa diffractometer 9 so that<br />

angular ranges of up to 60° in Ψ and 50 ° in β<br />

can be collected by merging several 6-8° strips,<br />

collected with the linear detector 11 , at a time.<br />

An even more subtle kind of organization was<br />

found for Ru-bpy molecules deposited onto<br />

ITO in device-like samples. Ru bpy, a Ru ++ -<br />

bipyridine 3<br />

complex with two PF - counterions,<br />

is known as an ionic conductor which produces<br />

efficient red light emission in organic lightemitting<br />

devices (OLED). In previous work it had<br />

been concluded that such Ru-bpy films were<br />

amorphous, as no discernable scattering was<br />

obtained with lab-based x-ray generators.<br />

Using GI-RSM we found subtle broad ring-like<br />

scattering features, after we optimized the<br />

scattering geometry to suppress scattering<br />

from the substrate. We could correlate the<br />

weak powder rings with a known crystalline<br />

structure and concluded that Ru-bpy forms<br />

nanoscale domains of only a couple of lattice<br />

constants in diameter. In addition we traced<br />

back our observation, that occasionally films<br />

featured higher order, to water contamination<br />

in the glove box. This first proof of GI-RSM<br />

being able to reveal the medium range order<br />

in a nominally amorphous organic film was<br />

recognized in an inside cover of Journal of<br />

Materials Chemistry 12 and Daniel Blasini was<br />

awarded the CHESS Thesis Prize for this work<br />

and his other results at G2 station 13 .<br />

Page 46 CHESS News Magazine 2009<br />

Fig. 3: GI-RSM of<br />

thermally grown<br />

pentacene on<br />

a silicon wafer.<br />

This data set was<br />

used in Nabok et<br />

al. 10 to identify<br />

the structure of<br />

the pentacene<br />

thin film phase<br />

in comparison<br />

to first principles<br />

calculations.<br />

Fig. 4: Ru-bpy<br />

based-light<br />

emitting device<br />

and Ru bpy<br />

molecule depicted<br />

on top of the<br />

scattering map<br />

obtained at G2 12 .<br />

Molecular monolayers are the ultimate<br />

challenge in surface diffraction. The Allara<br />

group at Penn State has developed a wetchemistry<br />

procedure of preparing thiol<br />

monolayers on GaAs surfaces. In Fouriertransform<br />

infrared (FTIR) spectra the thiols<br />

appeared well aligned, so the question was,<br />

whether these thiols would form well-ordered<br />

monolayers, as known for thiols on gold 14 .<br />

Christine McGuinness’ and coworkers’ paper<br />

in ACS Nano 15 ended a ten year quest of optimizing and characterizing the order in thiol layers on GaAs(100). It turned out that<br />

thiols on GaAs(100) only form nanoscale domains, even under the best preparation conditions. These nanoscale domains show<br />

some preferential alignment along the [100] substrate azimuth, which we ascribed to alignment of thiol crystallites parallel to step<br />

edges. Moreover, the lattice spacing in the thiol domains shows a subtle variation as a function of the substrate azimuth. This indepth<br />

study with about 30 samples became the most-cited paper in ACS-Nano in the first year of its introduction.<br />

As the flux and stability of the beam at G-line has continuously improved, we have been increasingly facing the problem that the<br />

limited dynamic range of the ORDELA linear gas detector has become the bottleneck that limits the speed of data acquisition,<br />

since intense peaks need to be rescanned with attenuation. A couple of years ago, we started talking to Peter Siddons at NSLS<br />

about the possibility of obtaining one of his prototype diode array detectors. In such an advanced detector, each diode element<br />

samples a comparable solid angle like the corresponding MCA channel of the gas detector. However, as each element has its<br />

own electronics, it can handle a dynamic range from single photon counting to a couple of 10 5 counts/sec, while the gas detector<br />

shows signs of saturation already at 10 3 counts/sec/channel.


At the time of writing we have just obtained such a high<br />

performance diode array and it is currently being set up at<br />

G2 by Abruña student Michael Lowe and G2 scientist Arthur<br />

Woll. State-of-the-art extended GI-RSMs, such as the one<br />

shown in Fig. 5, can thus be scanned in future without beam<br />

attenuation or fear of saturation of the brightest peaks.<br />

The new detector should help us to overcome many of the<br />

restrictions imposed by the limited dynamic range of the<br />

linear gas detector, promising a bright future for the GI-RSM<br />

system at G2.<br />

Acknowledgements<br />

I am indebted to many colleagues and coworkers, first of all<br />

Daniel Blasini, <strong>Cornell</strong> Chemistry, who built the reciprocal<br />

space mapping system at G2 with me. The list continues with<br />

Hector Abruña, <strong>Cornell</strong> Chemistry, Christine McGuinness and<br />

Fig. 5: Extended<br />

GI-RSM map of a<br />

pentacene film<br />

in the thin film<br />

phase. Several 8°<br />

strips, obtained<br />

by offsetting<br />

the delta arm of<br />

the G2 kappa<br />

diffractometer,<br />

were merged to a<br />

map spanning 40°<br />

in each direction of<br />

the detector arm 11 .<br />

David Allara, Penn State Chemistry, George Malliaras and<br />

Aram Amassian, <strong>Cornell</strong> Materials Science, Roland Resel and his merry men from the Technical <strong>University</strong> in Graz, Austria, Hisao<br />

Yanagi at Kobe <strong>University</strong>, Japan, Ed Kintzel, Jr., now at Western Kentucky <strong>University</strong>, as well as the other G2 students Dave Nowak,<br />

Yi Liu, and Aaron Vodnick. In the name of the G2 user community, we are most grateful to Pete Siddons, NSLS, for supplying us<br />

with a new diode array detector. Finally we thank the CHESS and G-line staff, in particular Arthur Woll and Ernie Fontes, for the<br />

good working conditions at G2 and all the technical help provided.<br />

References<br />

1. R. Feidenhans’l; “Surface Structure Determination by X-ray Diffraction”, Surf. Sci. Rep. 10, 105-188 (1988)<br />

2. I.K. Robinson, and D.J. Tweet; “Surface X-ray Diffraction”, Rep. Prog. Phys. 55, 599-651 (1992)<br />

3. J. Als-Nielsen, D. Jacquemain, D. Kjaer, F. Leveiller, and L. Leiserowitz; Phys. Rep. 246, 252 (1994); V. Kaganer, H. Mohwald, and<br />

P. Dutta; "Structure and Phase Transition in Langmuir Monolayers", Rev. Mod. Phys. 71, 779 (1999)<br />

4. D.-M. Smilgies, N. Boudet, B. Struth, Y. Yamada, and H. Yanagi; “<strong>High</strong>ly Oriented POPOP Film Grown on the KCl(001) Surface”, J.<br />

Cryst. Growth 220, 88-95 (2000); D.-M. Smilgies, N. Boudet, and H. Yanagi; “In-plane Alignment of Para-sexiphenyl Films Grown<br />

on KCl(001)”, Appl. Surf. Sci. 189, 24-30 (2002)<br />

5. S. Schiefer, M. Huth, A. Dobinevski, and B. Nickel; “Determination of the Crystal Structure of Substrate-induced Pentacene<br />

Polymorphs in Fiber Structured Thin Films”, J. Am. Chem. Soc. 129, 10316 (2007)<br />

6. A.C. Mayer, R. Ruiz, H. Zhou, R.L. Headrick, A. Kazimirov, and G.G. Malliaras; “Growth Dynamics of Pentacene Thin Films: Real-time<br />

synchrotron x-ray scattering study”, Phys. Rev. B. 73, 205307 (2006)<br />

7. D.-M. Smilgies, D.R. Blasini, S. Hotta, and H. Yanagi; “Reciprocal Space Mapping and Single-Crystal Scattering Rods”, J.<br />

<strong>Synchrotron</strong> Rad. 12, 807–811 (2005)<br />

8. D.-M. Smilgies, and D.R. Blasini; “Indexation Scheme for Oriented Molecular Thin Films Studied with Grazing-incidence Reciprocalspace<br />

Mapping”, J. Appl. Cryst. 40, 716-718 (2007)<br />

9. D.E. Nowak, D.R. Blasini, A.M. Vodnick, B. Blank, M.W. Tate, A. Deyhim, D.-M. Smilgies, H. Abruña, S.M. Gruner, and S.P. Baker;<br />

“Six-circle Diffractometer with Atmosphere- and Temperature-controlled Sample Stage and Area and Line Detectors for use in the<br />

G2 Experimental Station at CHESS”, Rev. Sci. Instrum. 77, 113301 (2006)<br />

10. D. Nabok, P. Puschnig, C. Ambrosch-Draxl, O. Werzer, R. Resel, and D.-M. Smilgies; “Crystal and Electronic Structures of Pentacene<br />

Thin Film from Grazing-incidence X-ray Diffraction and First-principles Calculations”, Phys. Rev. B. 76, 235322 (2007)<br />

11. D.-M. Smilgies, A. Amassian, S. Hong, S. Bhargava, J.R. Engstrom, and G.G. Malliaras; unpublished.<br />

12. D.R. Blasini, J. Rivnay, D.-M. Smilgies, J.D. Slinker, S. Flores-Torres, H.D. Abruña, and G.G. Malliaras; “Observation of Intermediaterange<br />

Order in a Nominally Amorphous Molecular Semiconductor Film”, J. Mater. Chem. (Hot Commun.) 17, 1458–1461 (2007)<br />

13. D. Blasini; “Dan’s Big Adventure”, CHESS News Magazine, p. 39 (2005)<br />

14. P. Fenter, P. Eisenberger, and K.S. Liang; “Chain-length Dependence of the Structures and Phases of Ch3(Ch2)N-1sh Self-assembled<br />

on Au(111)”, Phys. Rev. Lett. 70, 2447-2450 (1993)<br />

15. C.L. McGuiness, D. Blasini, J.P. Masejewski, S. Uppili, O.M. Cabarcos, D.-M. Smilgies, and D.L. Allara; “Molecular Self-Assembly at<br />

Bare Semiconductor Surfaces: Characterization of a Homologous Series of n-Alkanethiolate Monolayers on GaAs(001)”, ACS-Nano<br />

1, 30–49 (2007)<br />

CHESS News Magazine 2009 Page 47


Self-Assembled Nano-Checkerboard Thin Films<br />

Studied by Reciprocal Space Mapping at CHESS<br />

Sean M. O'Malley 1 , Peter Bonanno 1 , Keun hyuk Ahn 1 , Andrei A. Sirenko 1 ,<br />

Alex Kazimirov 2 , Soon-Yong Park 3 , Yoichi Horibe 3 , and Sang-Wook Cheong 3<br />

1<br />

Dept. of Physics, New Jersey Institute of Technology<br />

2<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

Rutgers Center for Emergent Materials and Dept. of Physics and Astronomy, Rutgers <strong>University</strong><br />

Material systems that can self-assemble into nano-structures are of great interest because of their ability to form<br />

compositionally and spatially distinct regions otherwise unattainable through traditional materials processing techniques<br />

such as optical or electron beam lithography. The spinel oxide ZnMnGaO 4<br />

(ZMGO) was recently discovered to undergo a selfassembly<br />

process during slow annealing of bulk samples 1 . Nano-structure formation in ZMGO occurs during the annealing<br />

process and is driven by a combination of spinoidal decomposition between Jahn-Teller-active and -inactive ions and spatial<br />

separation through diffusion processes 1-4 . The bulk ZMGO crystals end up comprising Mn-rich and Mn-poor nanorods<br />

forming a checkerboard (CB) pattern within the a-b plane that is accompanied by a herringbone (HB) arrangement along the<br />

c-direction.<br />

For the purpose of future device fabrication, nano-scale structures should be available in thin-film form, whereby the material<br />

properties can be enhanced through the overall reduction in volume and by the strain associated with heteroepitaxial<br />

growth. Our ZMGO thin films were grown using pulsed laser deposition by the research team at Rutgers <strong>University</strong> Center for<br />

Emergent Materials 5 . The target material was a homogenous high-temperature quenched form of ZMGO, with a tetragonal<br />

lattice (a = 0.82 nm, c = 0.87 nm); the substrate was single crystal cubic MgO (0 0 1) with lattice constant a = 0.4212 nm.<br />

Transmission electron microscopy (TEM) revealed that the films indeed contain an in-plane CB pattern with suppression of<br />

the out-of-plane HB formation. The typical nanorod dimensions in the epitaxial films increased to 4×4×700 nm 3 compared<br />

with 4×4×70 nm 3 for the bulk material. Portions of this work have been published in References 5,6 .<br />

Experiment<br />

To determine structural parameters and strain-accommodating distortions for the CB<br />

films we utilized reciprocal space mapping at the A2 x-ray diffraction (XRD) beamline<br />

of <strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong> (CHESS). The incident x-ray beam was<br />

conditioned using a double-bounce Si (1 1 1) monochromator passing photons with<br />

an energy of 10.531 keV (λ=1.1775 Å). <strong>High</strong> angular resolution was achieved by using<br />

a single-bounce Si (1 1 1) analyzer crystal. Reciprocal space mapping of the ZMGO<br />

checkerboard structures were measured for several symmetric and asymmetric<br />

reflections: (0 0 2), (−2 0 2), (−2 −2 2), (0 0 4), (0 4 4), (−2 2 6). The Miller index L<br />

corresponds to the film growth direction, while H and K represent the in-plane<br />

parameters of the structure. The integer values of H, K, and L correspond to reciprocal<br />

lattice points (RLP) of a cubic spinel structure with lattice parameter a S<br />

= 0.8424 nm<br />

(i.e., twice the corresponding value for the cubic MgO substrate).<br />

Results<br />

Figure 1 depicts a 3-dimensional construct of<br />

H−K, H−L, and K−L cross sectional maps measured<br />

around the asymmetric (0 4 4) RLP. The map is<br />

dominated by several peaks, associated with<br />

different phases within the CB film. The four broad<br />

peaks labeled α, ß, γ, and δ correspond to two<br />

conversely rotated tetragonal (ß and γ) and two<br />

perpendicularly-oriented orthorhombic (α and<br />

δ) phases, respectively. The four diagonal peaks<br />

(ρ, σ, τ, υ) correspond to the structural distortions<br />

originating from domain boundaries between<br />

checkerboard phases. The average in-plane<br />

lattice parameter of the two rotated tetragonal<br />

phases are lattice-matched to the substrate, i.e.<br />

the tetragonal phase is elastically strained with<br />

in-plane lattice constant of a tet<br />

= 0.841<br />

nm. These tetragonal phases of the<br />

ß and γ peaks are rotated by ±2.55°<br />

around the (0 0 L) reciprocal lattice<br />

vector. The orthorhombic phases α and<br />

δ have short and long in-plane lattice<br />

parameters a otho<br />

= 0.814 nm and b otho<br />

=<br />

0.898 nm, respectively, and are therefore<br />

inelastically strained. The diffraction<br />

peak intensities of the structural phases<br />

were integrated over their volume in<br />

reciprocal space in order to determine<br />

the prevalence of each phase within<br />

the film. The total intensity of the four<br />

α, ß, γ, and δ peaks is ~6 times that of<br />

the central tetragonal phase (A). This<br />

Fig. 1: (a) The H−K, H−L, and K−L cross-sectional<br />

RSM measured around the asymmetric (0 4<br />

4) reflection, with L = 4.08. The α and δ peaks<br />

are orthorhombic phases, while the ß and γ<br />

peaks are rotated tetragonal phases associated<br />

with the CB structure. The elastically strained<br />

tetragonal phase is labeled A. Signal originating<br />

from domain boundaries, highlighted by the<br />

radial lines emanating from the central (0 4<br />

4) RLP, are labeled ρ, σ, τ and υ. The central<br />

figure is a 3D reconstruction of experimentally<br />

determined FWHMs of the corresponding peaks.<br />

Reprinted figure with permission from Reference<br />

[6]. Copyright (2008) by the American Physical<br />

Society. (b) TEM image of the CB film on MgO<br />

substrate.<br />

Page 48 CHESS News Magazine 2009


atio is consistent with results from TEM<br />

images, and supports the assumption<br />

that this central peak A originates from<br />

the elastically strained transition layer<br />

between the CB layer and MgO substrate.<br />

Discussion<br />

The CB epilayer therefore contains four<br />

structurally different spinel phases: two<br />

conversely rotated tetragonal (Mnpoor,<br />

JT inactive) and two orthogonally<br />

oriented orthorhombic (Mn-rich, JTactive)<br />

phases. Each of the 4×4×750<br />

nm 3 nano-rod domains is formed<br />

by one of these four distorted<br />

unit cell phases as depicted in<br />

Fig. 2(a). The dashed lines in<br />

Fig. 2(b) define the CB supercell,<br />

which through translational<br />

operations can produce the entire<br />

CB film. The orthorhombic and<br />

rotated tetragonal phases are separated<br />

by domain boundaries (DB) closely<br />

aligned along the and <br />

directions. These domain walls should<br />

accommodate structural distortions<br />

between the α, ß, γ, and δ phases and<br />

provide a means for their coherent growth<br />

along both the film growth direction and<br />

in the a−b plane. Comparing the in-plane<br />

domain size (4×4 nm 2 ) and the in-plane<br />

footprint of the spinel lattice (~ 0.8×0.8<br />

nm 2 ), we determined that the fraction of<br />

distorted unit cells at the DB is significant<br />

(~30%) in comparison to the number of<br />

undistorted cells within each phase, see<br />

Fig. 2(b). Hence, these DBs should also<br />

produce a significant contribution to the<br />

total diffraction picture as illustrated by<br />

the prominent diagonal streak in the (0 4<br />

4) RSM.<br />

The symmetry of the lattice distortions at<br />

the domain boundaries of the CB pattern<br />

Fig. 2: (a) Illustration of the four structural<br />

unit cell phases present within the CB layer<br />

with arrows pointing to the appearance with<br />

the asymmetric (0 4 4) RSM. (b) Illustration<br />

of the CB arrangement of tetragonal and<br />

orthorhombic domains with DB along the<br />

and directions. The dashed<br />

lines define the CB super-cell (SC) by which<br />

translational operations can repeat the<br />

entire CB film.<br />

can be described in terms of distortions<br />

with respect to a 2D square lattice with<br />

monoatomic basis 7 . These distortions can<br />

be represented as a linear combination of<br />

e 3<br />

, r, and two other modes, either t +<br />

and<br />

s +<br />

or t –<br />

and s –<br />

. For example, distortions<br />

at the interface between the γ and δ<br />

domains, as illustrated in Fig. 3, can be<br />

characterized by a combination of the<br />

–e 3<br />

+ r + t +<br />

+ s +<br />

distortion modes with<br />

amplitudes of e 3<br />

= –ε/2, r = ε/2, t +<br />

= ε/2,<br />

and s +<br />

= ε/2.<br />

Fig. 3: Representation of the linear combination<br />

of distortion modes at the domain boundary τ.<br />

Squares γ and rectangles δ correspond to the rotated<br />

tetragonal and orthorhombic lattices, respectively.<br />

The role that the substrate plays in<br />

establishing and stabilizing the CB<br />

structure should not be overlooked as<br />

its influence can be seen in a number<br />

of aspects, e.g., the suppression of the<br />

out-of-plane herringbone formation.<br />

In the case of ZMGO thin films grown<br />

on MgAl 2<br />

O 3<br />

substrates (a S<br />

= 0.8083 nm)<br />

the film is under a compressive strain<br />

in contrast to the previously described<br />

case of growth on MgO, where the<br />

film is under stabilizing tensile strain.<br />

Results from TEM and RSM reveal that<br />

ZMGO samples grown on MgAl 2<br />

O 3<br />

have<br />

regressed back to forming the out-plane<br />

herringbone structure. Evidence of the HB<br />

structure can be seen in Fig. 4(a), where<br />

the signal originates from a pair of tilted<br />

inelastic tetragonal phases and two tilted<br />

elastically strained tetragonal phases,<br />

which appears around the symmetric<br />

(0 0 4) reflection. The elastic tetragonal<br />

phases are accompanied by rotations<br />

around K and H axes by about ±0.9°<br />

and the inelastic tetragonal phases<br />

are tilted by about ±1.4°. The long<br />

diagonal strikes originate from the<br />

canted nanorod interfaces, similarly to<br />

the data shown in Fig. 1 for the CB<br />

films.<br />

Conclusion<br />

The use of synchrotron radiation-based<br />

reciprocal space mapping (RSM) was<br />

paramount in investigating the complex<br />

structural properties of epitaxially grown<br />

ZnMnGaO 4<br />

thin films on single crystal<br />

MgO (0 0 1) substrates. The ZnMnGaO 4<br />

films consist of a self-assembled<br />

CB structure of highly aligned and<br />

regularly spaced vertical nano-rods.<br />

Reciprocal space mapping revealed<br />

the strain-accommodating interaction<br />

between the phases of the CB structure<br />

through lowering of the volume strain<br />

energy, while maintaining long-range<br />

commensurability of the structure with<br />

the MgO substrate. Such a planar CB<br />

surface could be utilized as a template<br />

for the subsequent growth of<br />

ferromagnetic material and the<br />

possible construction of highdensity<br />

magnetic recording<br />

devices. Work at NJIT and Rutgers<br />

was supported by the DE-<br />

FG02-07ER46382 and the NSF-<br />

DMR-0546985.<br />

Fig. 4: (a) crosssectional<br />

H-L<br />

RSM taken of the<br />

symmetric (004)<br />

reflection from ZMGO<br />

thin film grown on<br />

MgAl 2<br />

O 3<br />

. (b) TEM<br />

image showing HB<br />

structure.<br />

References<br />

1. S. Yeo, Y. Horibe, S. Mori, C.M. Tseng, C.H.<br />

Chen, A.G. Khachaturyan, C.L. Zhang, and<br />

S-W. Cheong; App. Phys. Lett. 89, 233120<br />

(2006)<br />

2. C.L. Zhang, S. Yeo, Y. Horibe, Y.J. Choi, S.<br />

Guha, M. Croft, S-W. Cheong, and S. Mori;<br />

Appl. Phys. Lett. 90, 133123 (2007)<br />

3. Y. Ni, Y. M. Jin, and A. G. Khachaturyan;<br />

Acta Mat. 55, 4903 (2007)<br />

4. M.A. Ivanov, N. K. Tkachev, and A. Ya.<br />

Fishman; Low Temp. Phys. 25, 459 (1999)<br />

5. S. Park, Y. Horibe, T. Asada, L.S. Wielunski,<br />

N. Lee, P. L. Bonanno, S.M. O’Malley, A.A.<br />

Sirenko, A. Kazimirov, M. Tanimura, T.<br />

Gustafsson, and S-W. Cheong; Nano Lett.<br />

8, 720 (2008)<br />

6. S.M. O’Malley, P.L. Bonanno, K.H. Ahn,<br />

A.A. Sirenko, A. Kazimirov, M. Tanimura, T.<br />

Asada, S. Park, Y. Horibe, and S-W. Cheong;<br />

Phys. Rev. B 78, 165425 (2008), http://link.<br />

aps.org/abstract/PRB/v78/p165424<br />

7. K.H. Ahn, T. Lookman, A. Saxena, and A. R.<br />

Bishop; Phys. Rev. B 68, 092101 (2003)<br />

CHESS News Magazine 2009 Page 49


Structures from Solutions: Biomolecular<br />

Small-Angle Solution Scattering at MacCHESS<br />

Richard E. Gillilan<br />

Macromolecular Diffraction Division of<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Fig. 1: Scattering profiles of<br />

a lysozyme standard taken<br />

on both G1 and F2 stations.<br />

For approximately the<br />

same sample-to-detector<br />

distances and energies,<br />

G1 performs better in the<br />

low-angle range, while F2<br />

yields wider-angle data. The<br />

inset shows a typical protein<br />

solution scattering pattern<br />

and the triangular region<br />

used to produce the profile.<br />

Protein crystallography continues to<br />

be the prime source of high-resolution<br />

molecular structural information in<br />

biology. Its success depends totally<br />

upon the ability of proteins and other<br />

biomolecules to form well-ordered<br />

crystalline lattices under the right<br />

conditions. Many important biological<br />

molecules, however, are not well<br />

behaved in this regard, a fact that is<br />

becoming more apparent as biologists<br />

focus their investigations on new and<br />

more challenging systems. Molecular<br />

order is handy, but it isn't everything. As<br />

it turns out, there is valuable structural<br />

information to be found even in dilute<br />

solutions where proteins are randomly<br />

oriented.<br />

The two-dimensional scattering pattern<br />

of a solution is perfectly symmetrical<br />

but the intensity falls off in a Gaussian<br />

fashion from the beamstop outward.<br />

In practice, 2D images thus collected<br />

are integrated into one-dimensional<br />

scattering intensity profiles, I(q),<br />

parameterized by q = 4π sin(θ)/λ, with<br />

2θ being the scattering angle (see Figure<br />

1). The inset in Figure 1 shows a typical<br />

biological small-angle x-ray scattering<br />

(BioSAXS) image collected at CHESS's<br />

G1 line. Several important particlespecific<br />

parameters are derived from<br />

these scattering curves. The famous<br />

Guinier relation connects the width of<br />

the Gaussian falloff observed at small<br />

q to the radius of gyration Rg of the<br />

particle, a useful geometric parameter 1 .<br />

The absolute scattered intensity at<br />

the beamstop, I(0), is related directly<br />

to molecular mass. Particle volume<br />

and maximum diameter can also be<br />

obtained.<br />

Historically, much x-ray<br />

solution scattering work has<br />

involved comparison to models.<br />

Experimental scattering curves<br />

can be used very effectively to<br />

select from among different<br />

competing structural hypotheses<br />

by comparison with computed<br />

curves. Because solutions<br />

can be examined under near<br />

physiological conditions, it<br />

becomes possible to study such<br />

phenomena as concentrationdependent<br />

changes in<br />

oligomeric state as well as<br />

large conformational changes. Even<br />

structurally-disordered systems give<br />

characteristic scattering profiles;<br />

consequently much use of x-ray solution<br />

scattering has been made in the study<br />

of protein unfolding 2 .<br />

Just how much information solution<br />

scattering yields is still open to debate,<br />

but modern advances in synchrotron<br />

technology and computer algorithms have<br />

produced some very surprising results<br />

and the popularity of BioSAXS has grown<br />

dramatically in recent years as a result. Due<br />

largely to the software tools developed by<br />

Sohyun (Sarah) Kim, is now a senior<br />

at Ithaca <strong>High</strong> School, Ithaca NY. She<br />

started working for Richard Gillilan<br />

during the Summer of 2007 writing<br />

a Java program to analyze 100,000<br />

digital images of crystal growth. This<br />

past summer, she gave a poster on her<br />

work at the CHESS user meeting<br />

(pictured above). Currently, she is<br />

volunteering in MacCHESS's lab<br />

over in the Biotech building located<br />

on the <strong>Cornell</strong> <strong>University</strong> campus,<br />

growing protein crystals for use by<br />

MacCHESS staff in beamline testing.<br />

One set of crystals she grew were used<br />

in a recent x-ray microbeam study by<br />

MacCHESS that is currently under<br />

review for publication. Sarah is also<br />

now working on a project to isolate<br />

purple membranes from a species of<br />

salt loving bacteria as a standard for<br />

several MacCHESS experiments.<br />

Page 50 CHESS News Magazine 2009


Svergun and coworkers (www.emblhamburg.de/ExternalInfo/Research/<br />

Sax/software.html), reconstruction of<br />

actual low-resolution biomolecular<br />

shapes from simple scattering profiles<br />

has become routine. But how is<br />

it possible to derive a meaningful<br />

three-dimensional structure from an<br />

orientationally-averaged scattering<br />

pattern that is inherently onedimensional<br />

The answer lies in how additional<br />

constraints are applied to help limit the<br />

number of potential solutions. Consider,<br />

for example, the widely-used GASBOR<br />

method of reconstruction 3 . Intensity is<br />

modeled by applying the Debye formula<br />

to a set of N "dummy residues" and firstlayer<br />

solvent atoms:<br />

Each g i<br />

(q) can be either a form factor<br />

for an average amino acid residue (a<br />

"dummy residue"), or the form factor<br />

for a solvent atom. The r ij<br />

are distances<br />

between dummy residues. Good<br />

configurations of dummy residues<br />

should minimize the mean square<br />

difference χ 2 between calculated and<br />

measured intensities:<br />

The sum runs over the number of<br />

sample points m in the scattering profile<br />

and σ 2 is the standard deviation of the<br />

measurements at each point. † But the<br />

problem is that there are too many<br />

nonphysical arrangements of<br />

dummy residues that<br />

minimize χ 2 . Additional constraints must<br />

be considered to limit the solution space.<br />

GASBOR defines a penalty function P(r)<br />

that rejects unphysical configurations<br />

of residues by insisting that models<br />

have a realistic distribution of nearest<br />

neighbors, that they be connected<br />

regions, and that the center of mass be<br />

close to the origin. GASBOR seeks to<br />

minimize the goal function E = χ 2 +αP, for<br />

some fixed positive weight α. Random<br />

movements of residues are accepted<br />

or rejected based on their E values<br />

according to a simulated annealing<br />

protocol.<br />

Typically 10 or more independent<br />

simulated annealing calculations<br />

are performed so that results can be<br />

compared. The degree of consensus<br />

between different solutions is an<br />

indicator of confidence. GASBOR<br />

attempts to incorporate wide<br />

angle data and is more effective<br />

at modeling structural detail.<br />

For very large complexes,<br />

the programs DAMMIN and<br />

DAMMIF, which use a related<br />

"dummy atom" approach, are<br />

more appropriate 4 .<br />

Volkov and Svergun have<br />

reconstructed a variety of<br />

ideal model shapes such as<br />

rods, disks, hollow spheres,<br />

and cylinders as a means of<br />

validation for these methods,<br />

but this classic paper on the<br />

subject also serves as a useful guide<br />

to the kinds of intensity profiles to<br />

expect and what kind of fidelity can be<br />

achieved in reconstructions 5 . A typical<br />

low-resolution shape reconstruction is<br />

shown in Figure 2 as a transparent<br />

surface superimposed<br />

on a crystal structure of a zinc transport<br />

protein solved by Cherezov et al. 6 .<br />

BioSAXS at CHESS<br />

While SAXS measurements are done at<br />

a number of CHESS beamlines, several<br />

time slots on G1 station are regularly<br />

dedicated to BioSAXS during each run.<br />

Typically running at 9-10 keV for SAXS,<br />

the multilayer optics feeding G1 deliver<br />

a very high flux with approximately 1%<br />

bandwidth. Disposable sample cells,<br />

custom manufactured by ALine Inc.<br />

(Redondo Beach, CA) require only 10 µl<br />

of sample and are produced with a layer<br />

of pressure-sensitive adhesive on the<br />

outside to accommodate a wide variety<br />

of (adhesive-free) window materials such<br />

as mica or ultra-thin polyimide (Fig. 3). A<br />

temperature-regulated housing for the<br />

sample cells has been fabricated which<br />

allows easy access for refilling.<br />

Fig. 3: Disposable sample cell for biological<br />

small-angle scattering (right). On-cell<br />

adhesive allows a wide variety of choices<br />

for thin-window material. The copper<br />

sample cell mount (left) can be easily<br />

removed from the temperature-controlled<br />

housing for refill.<br />

Fig. 2: Low resolution molecular shape<br />

reconstruction (transparent blue surface)<br />

compared with one of the known<br />

structures of the zinc transport protein<br />

CzrB from Thermus thermophilus.<br />

†<br />

For simplicity, scaling and correction factors have been omitted.<br />

Interested readers should consult Ref [3] for details.<br />

CHESS News Magazine 2009 Page 51


The beamstop in a SAXS<br />

experiment is placed as far<br />

from the sample as possible<br />

in order to capture scattering<br />

at the smallest angles. This<br />

necessitates using a vacuum<br />

or helium flight path to<br />

eliminate air scatter from<br />

the direct beam that would<br />

otherwise drown out the faint<br />

SAXS signals. The beamline<br />

configuration for BioSAXS is<br />

shown in Figure 4. G1 utilizes<br />

interchangeable vacuum pipes<br />

of various lengths containing<br />

an internally-mounted<br />

beamstop with integral PIN<br />

diode for monitoring direct<br />

beam counts. Sample-to-detector distances<br />

on the order of 850 mm are commonly used<br />

for proteins, yielding a practical q range from<br />

about 0.02 to 0.4 (300 Å - 15 Å), though G1 can<br />

accommodate significantly longer distances if<br />

necessary for large complexes.<br />

The Shannon sampling theorem holds that<br />

SAXS scattering profiles are fully determined<br />

by a relatively small number of parameters † ;<br />

consequently detector resolution needs are<br />

modest (i.e. the curves are gradual and have<br />

no real sharp features) 1 . Because additional<br />

physical constraints are always imposed when<br />

solving for molecular shapes (as illustrated<br />

above in the case of GASBOR), there is actually<br />

more information to be extracted from the<br />

data than one might expect on the basis of<br />

the sampling theorem. Nonetheless, 1k x 1k<br />

CCD detectors are usually more than adequate<br />

for BioSAXS needs. CHESS's unique FLICAM<br />

detector is routinely used for this purpose.<br />

While the number of pixels on state-of-theart<br />

crystallography detectors far exceeds the<br />

requirements for SAXS, they have been used<br />

quite effectively for BioSAXS on other stations<br />

and offer a potentially wider angle range for a<br />

single experiment.<br />

CHESS F2 station is now available for<br />

BioSAXS. F2's gentler flux, easy tunability,<br />

and high energy resolution are actually very<br />

well-suited for SAXS work. The increased<br />

flexibility in scheduling and access means<br />

that crystallography users can potentially<br />

collect BioSAXS and crystallographic data<br />

during the same visit. The same sample<br />

cell and temperature-control enclosure<br />

used in G1 can be used in F2; consequently<br />

there is no difference in sample handling.<br />

Fig. 4: BioSAXS beamline configuration at G1 station. Beam is collimated by<br />

a series of slits (right) and passes into the cooled sample cell enclosure. Both<br />

scattered radiation and direct beam travel through the vacuum flight tube<br />

where direct beam is stopped just prior to the detector. Scattering profiles<br />

are recorded on the FLICAM CCD detector at left.<br />

For convenience, a helium-filled flight path has been used in F2 with beam<br />

provided by a standard 300 µm collimator. Comparisons between G1 and F2<br />

using a lysozyme standard show excellent agreement (Figure 1). G1 reaches<br />

smaller angles q min<br />

and gives slightly better results in that region than F2, while<br />

F2 captures significantly wider angles q max<br />

than G1. Only a single quadrant of<br />

the Quantum 210 detector in F2 was used for these tests. Future improvements<br />

in beam collimation and beamstop design as well as planned extensions of<br />

sample-to-detector distance may well close the low-angle gap between the<br />

stations for most proteins, but G1 will remain the station of choice for the very<br />

largest complexes.<br />

Preparing for a run<br />

While BioSAXS is probably more tolerant of impurity than protein<br />

crystallography, good sample preparation is essential to success. Protein<br />

solutions should be monodisperse and as concentrated as possible. A total<br />

sample volume of 50 µl is the minimum advisable volume to prepare. This will<br />

allow enough for a series of dilutions, tests to determine the best exposure time,<br />

and any possible sample loading problems. Lysozyme gives an ideal signal at<br />

25mg/ml, but still gives a reasonable small-angle signal at 1 mg/ml. For much<br />

larger proteins, lower concentrations may be permissible since larger particles<br />

scatter more strongly. Concentrations of lysozyme stronger than 25 mg/ml<br />

will exhibit concentration-related distortions of the small-angle part of the<br />

scattering curve. It is advisable to collect data at several different concentrations<br />

and extrapolate to infinite dilution if possible. Often, as in the case of 10-25mg/<br />

ml lysozyme, a single dilute measurement is enough for good data, but<br />

concentration effects are heavily sample dependent and should be checked in<br />

each case.<br />

Users must also provide a matched buffer solution for background subtraction.<br />

This should be the exact same buffer used in the original sample preparation if<br />

possible. It is good to change buffer if you don't know the original composition<br />

exactly. Prepare plenty of extra buffer for sample dilutions and for rinsing<br />

sample cells.<br />

Users often wonder how many samples to bring and how long sample<br />

collection takes. Typically exposure times on G1 can be as short as one second<br />

(even with attenuation). Exposure times on F2 are commonly 20 sec or more.<br />

†<br />

Theoretically there are D max<br />

(q max<br />

-q min<br />

)/π parameters, where D max<br />

is the maximum diameter of the protein. By this criterion,<br />

measuring profiles down to q min<br />

< π/ D max<br />

is recommended. Often, the criterion q min<br />

< 1.3 ⁄ Rg is used in direct Guinier analysis.<br />

Page 52 CHESS News Magazine 2009


A complete dataset, however, may involve the collection of several<br />

protein concentrations along with matched buffers at several<br />

different exposure times. Most users need less than 24 hours to<br />

collect their data. New users are advised to bring only 2-6 proteins<br />

at first, preferably some with known structures that can be used<br />

to judge how well SAXS will work on their particular system.<br />

Experienced users have been able to collect data on dozens of<br />

proteins during that time, generating several gigabytes of images.<br />

Each protein responds differently to radiation damage. Preliminary<br />

data quality is assessed by looking at the flatness of the buffer<br />

subtraction and the linearity of Guinier plots. While MacCHESS does<br />

not yet offer microfluidic sample cells, Hong and Hao have recently<br />

demonstrated that scanning the x-ray beam within sample cells can<br />

significantly reduce radiation damage 7 .<br />

While BioSAXS can make definitive statements about disordered<br />

systems, potential users should be advised that shape reconstruction<br />

is only as good as the degree of order in the sample. Samples<br />

existing in more than one conformational or oligomeric state require<br />

special treatment and may prove difficult to process.<br />

Preliminary processing can and should be performed onsite during<br />

data collection. Lower-resolution shape reconstructions using<br />

the recently released DAMMIF program 4 are quite fast, but final<br />

reconstructions may require multiple independent calculations<br />

using slower algorithms, such as GASBOR 3 , that can take one or more<br />

days. Users can take advantage of MacCHESS's fast multiprocessor<br />

computers, cf105 and Feynman, to run processing jobs on site.<br />

Because BioSAXS involves the measurement of relatively few<br />

parameters to reconstruct complex shapes, and because it is<br />

sensitive to multiple complicating factors, such as aggregation,<br />

concentration effects, non-uniformity of sample cells and so forth,<br />

it is important to use it in conjunction with other techniques. In<br />

BioSAXS, much more so than in crystallography, it is important to<br />

recognize when data are not good and to utilize known standards<br />

when possible to assess the success of data collection and the<br />

fidelity of reconstructions.<br />

BioSAXS offers a comparatively simple and rapid means of learning<br />

about the size, low-resolution shape and oligomeric state of<br />

biomolecular complexes under near-physiological conditions. It has<br />

become an important tool in the reconstruction of large complexes<br />

for which only some parts are known. A number of groups have<br />

recently taken advantage of BioSAXS data collected on G1 station to<br />

complement their crystallographic studies and to resolve important<br />

questions on solution-state behavior. Zoltowski et al. investigated<br />

conformational switching in the fungal light sensor VIVID and were<br />

able to observe conformational changes between light and dark<br />

states in solution 8 . Wedekind et al. produced a model of catalyticallyactive<br />

APOBEC3G (a cytidine deaminase active on HIV singlestranded<br />

DNA) using BioSAXS data which has been further validated<br />

by recent experimental work 9 . Wang et al. studied calcium sensing by<br />

GCaMP2, a protein used in studying Ca 2+ flux in vivo in the heart and<br />

vasculature of mice 10 . Yuan et al. were able to identify the dimeric<br />

solution state of a serine integrase catalytic domain 11 . Hong and<br />

Hao have recently used wide-angle solution scattering data to help<br />

phase crystal structures 12 .<br />

Users interested in trying BioSAXS may apply for time through the<br />

express-mode proposal mechanism by specifying "Other" under<br />

"Choice of experimental technique" and by typing "standard<br />

BioSAXS" in the box provided for "Special experimental<br />

and facility needs." More information is available on the<br />

MacCHESS web site www.macchess.cornell.edu under the<br />

BioSAXS link.<br />

References<br />

1. D.I. Svergun, and M.H.J. Koch; "Small-angle Scattering<br />

Studies of Biological Macromolecules in Solution",<br />

Reports on Progress in Physics 66(10): p. 1735-1782<br />

(2003)<br />

2. N. Ando, et al., "Structural and Thermodynamic<br />

Characterization of T4 Lysozyme Mutants and<br />

the Contribution of Internal Cavities to Pressure<br />

Denaturation", Biochemistry 47(42): p. 11097-11109<br />

(2008)<br />

3. D.I. Svergun, M.V. Petoukhov, and M.H.J. Koch;<br />

"Determination of Domain Structure of Proteins from<br />

X-ray Solution Scattering", Biophysical Journal 80(6): p.<br />

2946-2953 (2001)<br />

4. D. Franke, and D.I. Svergun; "DAMMIF, a program for<br />

rapid ab-initio shape determination in small-angle<br />

scattering", Journal of Applied Crystallography, 42(2)<br />

(2009)<br />

5. V.V. Volkov, and D.I. Svergun; "Uniqueness of ab initio<br />

Shape Determination in Small-angle Scattering",<br />

Journal of Applied Crystallography 36: p. 860-864<br />

(2003)<br />

6. V. Cherezov, et al.; "Insights into the Mode of Action<br />

of a Putative Zinc Transporter CzrB in Thermus<br />

Thermophilus", Structure 16(9): p. 1378-1388 (2008)<br />

7. X.G. Hong, and Q. Hao; "Measurements of Accurate<br />

X-ray Scattering Data of Protein Solutions using<br />

Small Stationary Sample Cells", Review of Scientific<br />

Instruments 80(1): (2009)<br />

8. B.D. Zoltowski, et al.; "Conformational Switching in<br />

the Fungal Light Sensor Vivid", Science 316(5827): p.<br />

1054-1057 (2007)<br />

9. R.P. Bennett, et al.; "APOBEC3G Subunits Self-associate<br />

via the C-terminal Deaminase Domain", Journal of<br />

Biological Chemistry 283(48): p. 33329-33336 (2008)<br />

10. Q. Wang, et al.; "Structural Basis for Calcium Sensing by<br />

GCaMP2", Structure 16(12): p. 1817-1827 (2008)<br />

11. P. Yuan, K. Gupta, and G.D. Van Duyne; "Tetrameric<br />

Structure of a Serine Integrase Catalytic Domain",<br />

Structure 16(8): p. 1275-1286 (2008)<br />

X. Hong, and Q. Hao;<br />

12. "Combining Solution Wide-angle<br />

X-ray Scattering and Crystallography: determination of<br />

molecular envelope and heavy-atom sites", Journal of<br />

Applied Crystallography, 42(2): p. 259-264 (2009)<br />

CHESS News Magazine 2009 Page 53


Development of X-ray Pixel Array Detectors<br />

Sol M. Gruner 1,2<br />

1<br />

Department of Physics, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

X-ray detector development has not<br />

kept pace with the exponentially<br />

increasing brightness and capability of<br />

synchrotron sources. The result is that<br />

experiments at synchrotron sources<br />

are very frequently more limited by<br />

the detector than the source. The<br />

reason detector development has not<br />

kept pace is because responsibility<br />

for detector development has<br />

traditionally been somewhere<br />

between the sources and the users:<br />

Detectors have most commonly<br />

been viewed as the responsibility<br />

of the user end of a facility. In<br />

consequence, the large resources that<br />

have been devoted to developing<br />

new synchrotron facilities have not<br />

involved adequate development of<br />

new detectors. But it takes the better<br />

part of a decade, a dedicated team<br />

of detector designers, and millions<br />

of dollars to develop a new detector<br />

technology. This requires resources<br />

and a commitment that are very<br />

unusual for user groups. At the same<br />

time, the market is sufficiently small,<br />

and the commitment of resources<br />

is sufficiently large, that it isn’t very<br />

attractive to industry.<br />

A CHESS goal has been to break<br />

this cycle in order to catalyze<br />

the introduction of new detector<br />

technologies for synchrotron<br />

experiments. In the mid-1980’s CHESS<br />

teamed with Kodak (Rochester,<br />

NY) to install a novel image plate<br />

detector system. In the early 1990’s<br />

CHESS teamed with my group, then<br />

at Princeton <strong>University</strong>, to install<br />

the first fiber-optically coupled CCD<br />

detectors for routine macromolecular<br />

crystallography. This evolved into<br />

a collaboration with Area Detector<br />

Systems Corp (ADSC; Poway, CA) and<br />

resulted in ADSC’s very successful line<br />

of CCD detectors. Since the mid 1990’s,<br />

my group, now at <strong>Cornell</strong>, has been<br />

developing x-ray Pixel Array Detectors<br />

(PADs), a next generation detector<br />

technology.<br />

In a phosphor-coupled CCD detector<br />

x-rays are absorbed in a thin phosphor<br />

layer resulting in visible light emission.<br />

The phosphor is deposited on fiber<br />

optics that conveys the light image<br />

to a CCD. This type of detector,<br />

while having many advantages, still<br />

suffers from a number of drawbacks:<br />

Phosphors are slow and have limited<br />

resolution, fiber optics distort the<br />

image, each x-ray results in a signal<br />

in the CCD that is barely above the<br />

intrinsic noise, the CCD read time<br />

is long, and the technology allows<br />

limited in-detector processing of<br />

information.<br />

PADs can overcome all these<br />

limitations. In this context, a PAD<br />

consists of two silicon integrated<br />

circuits (ICs) bonded together (Fig. 1).<br />

The front layer, called the detective<br />

layer, is divided into pixels and is<br />

sufficiently sensitive throughout<br />

its ~500 μm thickness to stop<br />

most x-rays below roughly 20 keV.<br />

Each pixel of the detective layer is<br />

individually connected to its own<br />

Diode Detection Layer<br />

• Fully depleted, high resistivity<br />

• Direct x-ray conversion in Si<br />

Connecting Bumps<br />

• Solder, 1 per pixel<br />

CMOS Layer<br />

• Signal processing<br />

• Signal storage & output<br />

pixel of processing electronics in<br />

the back, or CMOS layer, via a small<br />

solder connection, or bump. Modern<br />

IC fabrication methods allow an<br />

incredible amount of processing<br />

power to be squeezed into each pixel.<br />

In a well designed PAD signals are<br />

conveyed to the processing layer in<br />

nanoseconds with amplitudes well<br />

above inherent noise, and with no<br />

signal spread beyond neighboring<br />

pixels. The processing layer can be<br />

designed not only to process signals,<br />

but also to read the information out<br />

very quickly.<br />

PADs are of two general types: photon<br />

counters and integrators. Each type<br />

has advantages and disadvantages,<br />

depending on the experiment at<br />

hand. Photon counters process<br />

photons one at a time, generally<br />

taking a few tenths of a μs to do so.<br />

Consequently, photons that arrive at<br />

a pixel faster than the time required<br />

to process the signal from an x-ray<br />

suffer counting losses. In other words,<br />

photon counters are locally count-rate<br />

limited. At the same time, they can<br />

provide some level of photon energy<br />

X-rays<br />

Fig. 1: A PAD, here shown in an exploded view, consists of two silicon ICs bonded<br />

together, pixel-by-pixel, with solder connecting bumps.<br />

Page 54 CHESS News Magazine 2009


discrimination. Integrators sum the signal<br />

from arriving x-rays for later digitization and<br />

can, thus, avoid local count rate limitations,<br />

but they do not discriminate between<br />

photons of different energies. Our focus is<br />

on integrators, in part because there are<br />

competent efforts in Europe (such as the<br />

PILATUS PAD developed at the Swiss Light<br />

<strong>Source</strong> (now being vended by DECTRIS,<br />

Switzerland) focused on photon counters.<br />

Specifically, we have developed two distinct<br />

integrating PADs.<br />

The first is a PAD made in collaboration with<br />

ADSC. It is designed to deliver images at up<br />

to a 1 kHz rate, with no dead time and with<br />

an extraordinarily large dynamic range per<br />

pixel (~2 x 10 7 x-rays/pixel/msec). Design of<br />

this detector was the principal focus of Dan<br />

Schuette’s Ph.D. thesis when he was at <strong>Cornell</strong><br />

(Dan was one of the first recipients of a<br />

<strong>Cornell</strong> G-line Fellowship. He is now a scientist<br />

at Lincoln Laboratory in Massachusetts.)<br />

Figure 2 shows an x-radiograph of a Canadian<br />

dime taken with this PAD. Figure 3 shows the<br />

extraordinary dynamic range and sharp point<br />

spread function (e.g., per pixel resolution) of<br />

this PAD.<br />

Fig. 2:<br />

X-radiograph<br />

of a Canadian<br />

dime. The<br />

back was<br />

sanded off to<br />

thin the coin.<br />

Both PADs are now operational as single chip detectors about an inch<br />

across. Future work on both projects will be to make mosaics of chips<br />

to cover larger areas. In the meantime both PADs are being used for<br />

a variety of user x-ray experiments at CHESS and the APS, including<br />

coherent x-ray microscopy, time-resolved pulsed laser deposition<br />

growth of complex oxides, protein solution scattering, and timeresolved<br />

growth studies of carbon nanotube forests.<br />

The future of PADs is very bright, indeed.<br />

Fig. 3: (top) Diffraction<br />

from a large grain<br />

aluminum sample. The<br />

four panels are all of a<br />

single image taken at the<br />

CHESS F2 station at various<br />

display magnifications. No<br />

beam stop was used, which<br />

is why there is a large<br />

central peak. (bottom)<br />

A graph of intensity<br />

through the single image<br />

diffraction pattern. Note<br />

the logarithmic ordinate.<br />

Also note the small peaks<br />

about 10 pixels to the left<br />

and right of the central<br />

main beam peak. These peaks are 25,000 times weaker than the central peak<br />

and less than a mm removed from it. This type of extraordinary dynamic range<br />

and sharp point spread behavior is not possible with a phosphor coupled CCD<br />

detector. From W. Vernon, M. Allin, R. Hamlin, T. Hontz, D. Nguyen, F. Augustine,<br />

S.M. Gruner, Ng.H. Xuong, D.R. Schuette, M.W. Tate, and L J. Koerner; “First<br />

Results from the 128x128 Pixel Mixed-mode Si X-ray Detector Chip”, Proc. SPIE,<br />

Conference 6706, paper 29, U-1 to U-11 (2007).<br />

The second PAD is being designed for<br />

the coherent imaging station at the Linac<br />

Coherent Light <strong>Source</strong> (LCLS) at SLAC. X-rays<br />

from diffraction patterns from the LCLS<br />

arrive at the detector with the time structure<br />

of the x-ray laser, e.g., ~100 fs. It would be<br />

impossible to count photons at these rates.<br />

So the PAD was designed to have very good<br />

single photon sensitivity, but to integrate the<br />

signal. In other words, the PAD will clearly<br />

be able to see signals as small as 1 x-ray per<br />

pixel, yet can receive up to 3000 x-rays per<br />

pixel. The entire detector will frame at the<br />

LCLS repetition rate of 120 Hz. Figure 4 shows<br />

a radiograph of part of U.S. dollar bill taken in<br />

Cu Kα radiation. The contrast in the image is<br />

provided by the ink in the dollar bill. Pull out<br />

a dollar and look at the relevant feature to<br />

note the resolution.<br />

Fig. 4: Cu Kα radiograph of part of<br />

U.S. dollar bill. The display of the<br />

radiograph data has been adjusted<br />

to zoom in on the contrast variation<br />

coming from the green ink in the<br />

dollar.<br />

Acknowledgements:<br />

The PAD developments described in this article have been carried out by<br />

members of the <strong>Cornell</strong> detector group in my lab in the <strong>Cornell</strong> Physics<br />

Department, including Mark Tate, Hugh Philipp, Marianne Hromalik,<br />

Lucas Koerner, Kate Green and Dan Schuette (now at Lincoln Labs in<br />

Massachusetts). Darol Chamberlain (CHESS) has also been involved. PAD<br />

research in our lab is supported by DOE Office of Biological Research<br />

via grant DE-FG02-97ER62443. The ADSC project is a collaboration with<br />

Area Detector Systems Corporation via NIH grant RR-014613. LCLS PAD<br />

development is supported by the Department of <strong>Energy</strong>.<br />

CHESS News Magazine 2009 Page 55


Exploring New Physics of Nanoparticle Supercrystals<br />

by <strong>High</strong> Pressure Small Angle X-ray Diffraction<br />

Zhongwu Wang, Ken Finkelstein, and Detlef-M. Smilgies<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Superlattices built from self-assembled<br />

nanoparticles (NPs) combine the<br />

size-tuned unique properties of single<br />

crystals with the collective properties<br />

of newly created (one- two- and threedimensional)<br />

ordered arrays and promise<br />

the next generation of devices spanning<br />

a variety of applications 1-4 . For example,<br />

in nature, a colorless opal is a disordered<br />

aggregate of silica particles. When the<br />

silica particles are ordered, the opal<br />

changes to a color correlated with the<br />

size of the self-assembled particles 3 .<br />

In the laboratory, several approaches,<br />

including templating, soft lithography<br />

and supramolecular chemistry,<br />

have enabled creation of numerous<br />

superlattices comprising NP building<br />

blocks 3,5-7 . These NP nanobuilding blocks<br />

fall mostly into three types of order<br />

displaying variable packing densities<br />

and properties: Face-Centered-Cubic<br />

(FCC), Hexagonal-Centered-Packing<br />

(HCP) and Body-Centered-Cubic (BCC)<br />

or Body-Centered Tetragonal (BCT) (Fig.<br />

1) 4 . To facilitate custom engineered<br />

applications for NP superlattices<br />

requires understanding the structural<br />

and mechanical stabilities, inter-NP<br />

interaction and tuning mechanisms.<br />

In order to explore the structure and<br />

dynamic behavior at atomic resolution as<br />

NPs evolve into ordered building blocks<br />

and are tuned for stable structures and<br />

novel properties, we have developed<br />

a powerful in-situ synchrotron x-ray<br />

technique. Our window capable of<br />

peeking into the mysterious nanoworld<br />

utilizes small and wide angle x-ray<br />

diffraction under varying pressure and<br />

temperature for the exploration of<br />

the new physics and chemistry of NP<br />

supercrystals.<br />

Probing a wide spectrum of information<br />

of NP superlattice formation covering<br />

the atomic scale and the nanoscale<br />

requires monitoring x-ray diffraction<br />

from NP supercrystals in the small and<br />

the wide angle range. The complexity of<br />

measurement increases with application<br />

of pressure and involves several key<br />

experimental components: a pressure<br />

cell with transparent x-ray windows,<br />

x-ray energy tuning and position control<br />

of a large area detector. The diamond<br />

anvil cell (DAC) is certainly the primary<br />

option to reach extreme pressure, but<br />

strong absorption of low energy x-rays<br />

by diamond limits the measurement<br />

of small angle x-ray scattering. We<br />

circumvent these problems by tuning<br />

x-ray energy and sample-to-detector<br />

distance. At a wavelength optimized<br />

to avoid significant weakening of the<br />

scattering signal, we optimize detector<br />

position for collection of both firstorder<br />

small angle and wide angle x-ray<br />

diffraction by taking several exposures at<br />

different detector positions (but without<br />

multiple calibrations of sample-todetector<br />

distance).<br />

Our DAC has a large downstream<br />

angular opening that covers the wide<br />

angle x-ray diffraction. For a supercrystal<br />

sample made up of nanoparticles with<br />

known size, we roughly estimate the<br />

angle and corresponding position of the<br />

first order small angle x-ray diffraction<br />

peaks at different x-ray wavelengths,<br />

and then tune the x-ray energy to the<br />

optimum value at one intermediate<br />

sample-to-detector distance. After a<br />

quick check of the diffraction image, we<br />

move the detector to a long distance<br />

from the sample, enabling observation of<br />

small angle x-ray diffraction rings. After<br />

one calibration of sample-to-detector<br />

distance, we move the detector off the<br />

beam center. A one-dimensional plot<br />

generated from the collected x-ray<br />

diffraction pattern can then cover both<br />

small and wide angles, and processing<br />

of the two-dimensional pattern on the<br />

image requires only a simple definition<br />

of the beam center. Due to a significant<br />

reduction of the beam intensity at a low<br />

energy, the optimization of experimental<br />

conditions for collection of high quality<br />

images at short exposure times uses<br />

the following approach: the higher the<br />

x-ray energy, the smaller the beam stop<br />

and the shorter the sample-to-detector<br />

distance.<br />

The key component in this scheme is<br />

the development of a four dimensional<br />

control system for our Mar345 detector.<br />

As shown in Fig.2, tracks attached to<br />

the B2 hutch roof are used to move the<br />

detector, effectively eliminating detector<br />

vibration during data collection. Buttons<br />

2 and 3 control the movement of the<br />

detector horizontally, perpendicular<br />

and parallel to the beam, whereas<br />

knobs 4 and 5 control detector vertical<br />

position and rotation. At 25 keV, when<br />

the detector is ~450 mm from the<br />

sample, the small angle x-ray rings<br />

diffracted from a NP supercrystal are<br />

almost entirely blocked by the beam<br />

stop (Fig.2a). With x-ray energy tuned to<br />

20 keV, the first-order small angle x-ray<br />

Fig. 1: Schematic illustrating the<br />

positioning of nanobuilding blocks into<br />

periodic crystallographic arrays and<br />

corresponding small angle x-ray diffraction<br />

patterns. Top left shows a spherical<br />

nanobuilding block, top right a superlattice<br />

comprising nanobuilding blocks located<br />

at face centred cubic (FCC) lattice points.<br />

Adapted partially from Ref [4].<br />

Page 56 CHESS News Magazine 2009


diffraction rings are fully observable<br />

(Fig.2b), allowing refinement of the NP<br />

superstructure to BCT symmetry. In a<br />

FCC ordered supercrystal composed of<br />

slightly larger NPs, the full small angle<br />

x-ray diffraction pattern was observed by<br />

moving the detector distance to ~1200<br />

mm.<br />

Fig.3 shows one example of how well<br />

this technique worked for high pressure<br />

studies of NP supercrystals. By tuning<br />

the x-ray energy to several wavelengths<br />

at ambient conditions, the small angle<br />

and wide angle x-ray diffraction images<br />

of In 2<br />

O 3<br />

NP supercrystals 8 were collected<br />

and shown in Fig.3 (left). Small angle<br />

x-ray diffraction images (Figs.3a, 3b)<br />

confirm that 15 nm octahedral-shape<br />

NPs assemble into an ordered simple<br />

BCT superstructure. In addition, In 2<br />

O 3<br />

supercrystals display a noticeable<br />

lamellar structure that may be caused<br />

by a preferred orientation. Fig 3a was<br />

obtained using the small-angle setup<br />

at D1 with an energy of 10 keV and a<br />

sample to detector distance of 700 mm<br />

for comparison, and Fig 3b shows that<br />

our current set up at B2 can detect all<br />

SAXS rings with the exception of the first<br />

order.<br />

The wide angle x-ray diffraction image<br />

and integrated pattern (Fig.3c) indicate<br />

that each single In 2<br />

O 3<br />

NP crystallizes<br />

into a cubic atomic structure. A pressure<br />

induced hexagonal phase has been<br />

observed at 10 GPa in bulk materials,<br />

but it does not appear in this NP<br />

supercrystal. However, small angle x-ray<br />

diffraction (Fig.3d) clearly reveals that<br />

the intermediate distance between NPs<br />

expands at pressures above 10 GPa. This<br />

Fig. 2: (left) Four dimensional control of mar345<br />

detector: 1) Mar345 Detector; 2) side-to-side (x); (3)<br />

Back-forward (y); (4) Vertical (z); (5) Rotation (x-y);<br />

(6) Moving tracks. (Right) small angle images of<br />

supercrystal with same particle size: a) 25 keV for<br />

BCT at short distance; b) 20 keV for BCT at long<br />

distance; c) 20 keV for FCC with slightly larger NPs<br />

at long distance.<br />

enhanced structural stability could be<br />

attributed to combined effects of particle<br />

size, morphology and NP order; the<br />

pressure-induced expansion of inter-NP<br />

distance mostly results from a series<br />

of subtle surface structure variations.<br />

Additional examples have also been<br />

tested (Fig.1), and we observed several<br />

novel phenomena including: mechanical<br />

stability, phase transformation and<br />

associated changes in surface structure<br />

and nanoparticle interaction.<br />

This unique capability opens a window<br />

for exploring new phenomena and<br />

tuning mechanisms for different NPs and<br />

NP building blocks under pressure. The<br />

collected information enables not only<br />

understanding the mechanical stability,<br />

multiple interactions, processes and<br />

resulting mechanism of nanoparticles<br />

Fig. 3: Small and wide angle x-ray diffraction patterns of In 2<br />

O 3<br />

nanoparticle supercrystals at ambient and<br />

pressurized conditions. Panels a and b, small angle x-ray patterns at 0 pressure at different energies; c,<br />

wide angle x-ray image and integrated pattern; d, high pressure small angle x-ray scattering.<br />

and packed building blocks [9], but also<br />

clarifying the development of surface<br />

properties upon pressure-induced phase<br />

transformation in single nanoparticles.<br />

We appreciate technical assistance<br />

and discussions from all CHESS staff<br />

that made this high pressure technique<br />

feasible. Special thanks go to several<br />

collaborators including Prof. J. Fang<br />

(SUNY), Dr. H. Fan (SNL), Prof. D. C.<br />

Sayle (UK), Prof. T. Hyeon (Korea), Dr. R.<br />

Hoffmann (<strong>Cornell</strong>) for sample syntheses<br />

and valuable discussions.<br />

References:<br />

1. M.A. Meyers, A. Mishra,. and D.J.<br />

Benson; “Mechanical Properties of<br />

Nanocrystalline Materials”, Prog. Mater.<br />

Sci. 51, 427-556 (2006)<br />

2. A.P. Alivisatos; “Semiconductor Clusters,<br />

Nanocrystals, and Quantum Dots”,<br />

Science 271, 933-937 (1996)<br />

3. M.P. Pileni; “Self-assembly of Inorganic<br />

Nanocrystals: Fabrication and collective<br />

intrinsic properties”, Acc. Chem. Res. 40,<br />

685-693 (2007)<br />

4. D.C. Sayle, S. Seal, Z. Wang, et al;<br />

“Mapping Nanostructure: A systematic<br />

enumeration of nanomaterials by<br />

assembling nanobuilding blocks at<br />

crystallographic positions”, ACS Nano 2,<br />

1237-1251 (2008)<br />

5. B.D. Gates, Q.B. Xu, M. Stewart, et al;<br />

“New Approaches to Nanofabrication:<br />

molding, printing and other techniques”,<br />

Chem. Rev. 105, 1171-1196 (2005)<br />

6. Y.N.Xia and G.M. Whitesides; “Soft<br />

Lithography”, Ann. Rev. Mat. Sci. 28,<br />

153-184 (1998)<br />

7. J. Aizenberg, J.C. Weaver, M.S.<br />

Thanawala, V.C. Sundar, D.E. Morse,<br />

and P. Fratzl; “Skeleton of Euplectella sp.:<br />

Structural hierarchy from the nanoscale<br />

to the macroscale”, Science 309,<br />

275-278 (2005)<br />

8. W.Q. Lu, Q.S. Liu, Z.Y. Sun, J.B. He, C.<br />

Ezeolu, and J.Y. Fang; “Supercrystal<br />

Structure of Octahedral c-In 2<br />

O 3<br />

Nanocrystals”, J. Am. Chem. Soc. 130,<br />

6983-6991 (2008)<br />

9. D.K. Smith, B. Goodfellow, D.M.<br />

Smilgies, and B.A. Korgel; “Selfassembled<br />

Simple Hexagonal Ab2 Binary<br />

Nanocrystal Superlattices; SEM, GISAXS<br />

and Defects”, JACS, DOI: 10.1021/<br />

ja8085438, http://pubs.acs.org/doi/<br />

abs/10.1021/ja8085438<br />

CHESS News Magazine 2009 Page 57


Topography of Diamonds at CHESS<br />

Helps Nuclear Physics Program at JLAB<br />

Richard Jones 1 , Franz Klein 2 , and Ken Finkelstein 3<br />

1<br />

<strong>University</strong> of Connecticut<br />

2<br />

Catholic <strong>University</strong> of America<br />

3<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

Our group at the <strong>University</strong> of Connecticut, with<br />

collaborators from Catholic <strong>University</strong> of America and<br />

Glasgow <strong>University</strong>, is in charge of construction of the<br />

photon source for Hall D at Jefferson Laboratory (JLAB).<br />

This source of polarized high-energy photons is an<br />

essential part of the GlueX experiment [1] – a key scientific<br />

motivation for the 12 GeV Upgrade of CEBAF at JLAB and<br />

the highest priority in the 2007 NSAC Long Range Plan for<br />

Nuclear Science. GlueX requires a 9 GeV photon beam with<br />

a high degree of linear polarization that will be produced<br />

by passing 12 GeV electrons through a thin oriented<br />

diamond crystal radiator, using the process of coherent<br />

bremsstrahlung (CB).<br />

selection for diamond radiators. We were unable to<br />

identify any x-ray diffraction end station at major US<br />

facilities with the capability to do whole-crystal rocking<br />

curves on samples of dimensions 10 mm with arc-second<br />

resolution, but we were advised to contact CHESS to discuss<br />

our specialized requirements. CHESS staff member Ken<br />

Finkelstein agreed to collaborate with us on a feasibility<br />

study; first measurements of diamond radiators took place<br />

at CHESS in November 2006. During that run we obtained<br />

rocking curve images of several used diamond radiators.<br />

These whole-crystal rocking curves were taken with 100<br />

micron spatial resolution using a CCD detector - the first<br />

time these whole crystals had been examined with such<br />

high spatial resolution.<br />

Figure 2 illustrates our first topographic rocking curve<br />

setup that provided ~100 micron spatial resolution. The<br />

asymmetric Si(111) monochromator (green) operating at<br />

15KeV (b = 8) expanded the beam and gave sample (gold)<br />

rocking curve angle resolution not better than 150mrad<br />

(FWHM). Adding a second, symmetric Si(220) mono (pink)<br />

improved resolution to 30mrad. Photo shows sample & lenscoupled<br />

CCD mounted on C-line 4-circle diffractomator.<br />

Figure 3 displays transmission topographs of several<br />

diamond samples. Localized light regions near the center<br />

Fig. 1: Illistration of the relative position of the source<br />

of polarized photons (diamond wafer), GlueX target, and<br />

surrounding components making up the high energy<br />

physics detector. [image from www.gluex.org]<br />

The creation of CB makes very stringent demands on the<br />

thickness and mosaic spread of the diamonds. We start with<br />

the very best monocrystals available from the diamond<br />

industry and thin them to the required thickness while<br />

subjecting them to a thorough quality control process. The<br />

unprecedented intensity of the GlueX source demands a<br />

better understanding of radiation damage rates in diamond<br />

than is currently available. Because the “coherent scattering<br />

vertex” (scattering process) in CB has the same form as<br />

elastic scattering in x-ray diffraction, the later method is an<br />

essential tool for diamond radiator diagnostics.<br />

The ability to do whole-crystal rocking curves with arcsecond<br />

angular resolution is essential to raw material<br />

Figure 2<br />

Page 58 CHESS News Magazine 2009


in two images show damage caused by<br />

exposure to high energy electron beams at<br />

JLab. Near upper left corner are ink spots<br />

added by the manufacturer for identification.<br />

GlueX requires diamond radiators of thickness<br />

no more than 20 microns - the thinnest ever<br />

used in a CB source. The diamond we studied<br />

during this visit had been produced for use<br />

in Hall B, but never put into production<br />

because it appeared to be unstable in<br />

alignment. The measurements at CHESS<br />

provided an explanation for this instability:<br />

Figure 3<br />

the rocking curves demonstrated that the<br />

diamond was severely warped. This result was<br />

very significant because it pointed to a problem with the thinning and mounting techniques used for past CB sources.<br />

Uncovering this problem early in the R&D phase of the GlueX experiment has enabled us to put in place a mitigation plan<br />

that will help avoid project delays and additional costs that would result if not uncovered until beamline commissioning.<br />

Following this initial success, we returned for a second run in 2007. The purpose was primarily to help in the upgrade and<br />

testing of new C-line optics designed to obtain arc-second resolution in the rocking curves. This was an essential element<br />

of the original feasibility proposal, and would yield an order of magnitude improvement over what was obtained in 2006.<br />

The improvement was made possible by fabrication of a custom asymmetric silicon (311) mono crystal pair that provided<br />

nearly perfect dispersion matching to the diamond (220) reflection. The optics were designed and simulated by our<br />

group, and machined, etched & installed at C-line by CHESS specifically for our measurements. Subsequent measurements<br />

demonstrated the customization meets all requirements put forward for CB diamond radiator diagnostics.<br />

Figure 4, a photo taken in Nov. 2007, shows the new C-line asymmetric monochromator.<br />

It uses two asymmetric Si (311) reflections with b=14 at 15KeV, and provides a large<br />

beam that is dispersion matched to diamond 220 planes so dispersion broadening under<br />

10μrads was expected.<br />

Figure 5 shows results from a quantitative analysis of a series of topographs from a<br />

nearly perfect diamond. The indicated pixel range corresponds to a 3 square mm area.<br />

Vertical surface height variation shows the range of angle required for peak rocking curve<br />

intensity. The angular resolution obtained shows the new monochromator is working as<br />

expected.<br />

A follow-up run was conducted in May 2009 to take advantage of instrument upgrades<br />

at C-line (see p. 17) that enabled switchover to our custom monochromator more quickly<br />

and efficiently. In a related development, our group has benefited from<br />

parallel developments in thin diamond monocrystal fabrication and<br />

mounting taking place at the Instrumentation Division of Brookhaven<br />

National Lab. From this group, we obtained, on loan, a diamond<br />

monocrystal produced using the latest CVD diamond technology. We<br />

characterized the mosaic structure at CHESS during our trip in May.<br />

Figure 4<br />

Access to unique facilities at CHESS is essential to carry out diamond<br />

thinning and mounting studies for GlueX during the next three years.<br />

Once those goals have been met, we anticipate there will be an ongoing<br />

need for radiator diagnostics that will enable us to provide a continuous<br />

source of top-quality radiators to guarantee continuous operation<br />

of the source throughout the data-taking lifetime of GlueX. CHESS<br />

occupies a unique place among X-ray user facilities in the US as a place<br />

where specialized needs can be accommodated. As non-specialists<br />

in the techniques of X-ray optics and measurements, our group has<br />

greatly benefited and learned from our collaboration with CHESS staff.<br />

Figure 5<br />

Our experience demonstrates how CHESS has a critical impact in other areas of scientific exploration.<br />

[1] A goal of the GlueX experiment is to provide critical data needed to address one of the outstanding and fundamental challenges in physics – the<br />

quantitative understanding of the confinement of quarks and gluons in quantum chromodynamics (QCD). For more info see http://portal.gluex.org/.<br />

CHESS News Magazine 2009 Page 59


Complex Porous Metal Nanostructures:<br />

self-assembled from block copolymers<br />

Uli Wiesner 1 , Laura Houghton 2 , and Sol M. Gruner 2,3<br />

1<br />

Department of Materials Science & Engineering, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

Department of Physics, <strong>Cornell</strong> <strong>University</strong><br />

Uli Wiesner (Dept. of Materials Science &<br />

Engineering, <strong>Cornell</strong> <strong>University</strong>) and coworkers<br />

have pioneered a novel way to use block<br />

copolymers to make nanoporous structures<br />

for catalyst, battery and fuel cell applications.<br />

The self-assembly of nanoparticles using<br />

block copolymers depends largely on the<br />

ability to design the right polymers and<br />

ligand-stabilized nanoparticles. Ligands can<br />

be used to achieve a high solubility in an<br />

organic solvent and allow particles to flow<br />

at high density. The layer of ligand needs to<br />

be thin around each particle so that the final<br />

structure has a volume of metal large enough<br />

to maintain a microstructural shape when the<br />

organic materials are removed 1 . The metal<br />

particles may then be sintered or calcined into<br />

a metal or metal oxide framework that has<br />

structure directed by the parent copolymer<br />

system and high porosity.<br />

A structure made of platinum, which is<br />

the best catalyst for fuel cells, can now be<br />

created with uniform hexagonal pores. The<br />

neatly ordered pores create a large surface area within a very small volume of<br />

material. Fuel cells and solar cells also both benefit from nanoscale pores: For<br />

a fuel cell, pores offer increased surface area at which a fuel can interact with<br />

a catalyst, making it more efficient. Similarly in certain types of solar cells,<br />

more surface area means that more light can be absorbed and generated<br />

charge carriers can be collected, converting more of the incident light energy<br />

into electricity. Porous films of titanium oxide, used in Grätzel-type solar<br />

cells, and niobium oxide, a fuel cell catalyst support, are obtained by mixing<br />

the chemicals that react to form the metal oxides with a polymer solution of<br />

Pl-b-PEO (poly isoprene-block-polyethylene oxide). As the reaction proceeds,<br />

the Pl portion of the copolymer forms cylinders surrounded by metal oxides.<br />

Subsequent heat treatments leave uniform, crystalline metal oxide with<br />

cylindrical pores 2 . The exceptional properties make them excellent candidates<br />

for solar cells but also valuable in many other areas.<br />

In addition to making porous materials, this technique could be used to create<br />

finely structured surfaces. In the field of plasmonics, waves of electrons move<br />

across the surface of a conductor with the information-carrying capacity of<br />

fiber optics, but in spaces small enough to fit on a chip. The self-assembly of<br />

nanoparticles with block copolymers may enable the formation of porous<br />

materials made of a range of elements, alloys, or intermetallics. This may<br />

enable mixtures of distinct metal nanoparticles to be combined into a single<br />

nanoporous material. Nanoporous metals made from nanoscopic particles<br />

of distinct compositions may have unique electrical, optical, and catalytic<br />

properties. There is great promise for future energy<br />

applications. Control over the structure of metals is crucial<br />

for energy conversion, sensing, and information processing.<br />

Self-assembly of nanoparticles with block copolymers<br />

provides a natural entry point to materials structured on<br />

this length scale 3 .<br />

Related information by the Wiesner Research Group,<br />

Department of Materials Science at <strong>Cornell</strong> <strong>University</strong> can<br />

be found at http://people.ccmr.cornell.edu/~uli/. CHESS<br />

was used to characterize the nanoporous structures at<br />

various steps along the processing sequence in order to<br />

understand the formation process, thereby enabling future<br />

materials improvements.<br />

References:<br />

Computer simulated image of how platinum<br />

nanoparticles fuse into a structure with tiny pores<br />

after the polymers that guided them into position<br />

are removed. Figure from the Wiesner Lab.<br />

1. B.Steele; "In ‘novel playground’, metals are formed into porous<br />

nanostructures for better fuel cells and microchips", <strong>Cornell</strong><br />

Chronicle Online (June 26, 2006)<br />

2. B.Steele; "‘One-pot’ process can make more efficient materials<br />

for fuel cells and solar cells", <strong>Cornell</strong> Chronicle Online (January<br />

28, 2008)<br />

3. S.C. Warren, and U. Wiesner; “Self-assembled Ordered<br />

Mesoporous Metals”, Pure Appl. Chem 81(1), pp. 73-84 (2009)<br />

Page 60 CHESS News Magazine 2009


Heat-bump Measurements at CHESS A2 Wiggler Beam<br />

Peter Revesz 1 , Alex Kazimirov 1 , Ivan Bazarov 4 , Jim Savino 1 ,<br />

Emmett Windisch 2 , and Christopher MacGahan 3<br />

1<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

Wayne State <strong>University</strong><br />

3<br />

Applied Engineering and Physics, <strong>Cornell</strong> <strong>University</strong><br />

4<br />

Physics Department, <strong>Cornell</strong> <strong>University</strong><br />

Thermal distortion of the first crystal due to high heat load can significantly reduce the throughput of a<br />

monochromator. An in-situ optical technique has been developed to measure the resulting crystal distortion (heatbump).<br />

We have investigated the effect of the heat load from the CHESS A2 wiggler on crystals and multilayers. A<br />

computer algorithm has been developed to calculate the depth distribution of the absorbed power in a format that<br />

can be used by ANSYS. Calculations performed using this approach are in good agreement with the experimental<br />

profiles measured by the optical technique.<br />

Experimental Technique and Results<br />

The CHESS A-line 49 pole wiggler produces a total power of 6.6 kW when<br />

operating at 5.3 GeV and 200 mA current. Although only a portion of this<br />

power reaches the monochromator due to the apertures and upstream<br />

filters, still, the heat load can produce a considerable deformation of the<br />

first crystal, leading to the deterioration of the monochromatic beam<br />

(a)<br />

and the loss of x-ray flux throughput. Different approaches have been<br />

developed to deal with the heat load problem. They include internal<br />

and external cooling of Si crystals, including cryo-cooling, to minimize<br />

the thermal expansion 1 , using diamond instead of Si 2 and, in the case<br />

of multilayer optics, using SiC as a substrate material instead of Si 3 .<br />

Traditionally, the effect of the heat-bump is assessed by measuring<br />

a monochromator rocking curve. The widening of the rocking curve<br />

characterizes the deterioration of the Bragg reflection averaged over the<br />

beam footprint area.<br />

Our direct method of a heat-bump measurement gives an in-situ 3D map<br />

of the surface distortion. The main idea of the method proposed by P.<br />

Revesz et al. 4,5 is as follows: We use an array of light source dots viewed as<br />

(b)<br />

reflected from the monochromator crystal surface. As the surface heatbump<br />

is formed and the reflecting surface becomes distorted, the image measurement. The image of an array of light-dots is<br />

Fig. 1: (a) Principal scheme of the heat-bump<br />

of the regular dots pattern will also be distorted. More accurately, the<br />

captured with a CCD camera as reflected from the crystal.<br />

image of the individual dots will shift their positions according to the local The position of the dots in the image shifts as the crystal<br />

becomes deformed as a result of a beam heating. (b) Slope<br />

slope error developed on the crystal surface. The principal setup of the<br />

gradient field produced by a heat load and captured by<br />

experiment is shown in Figure 1a. The array of light dots was formed by<br />

a heat-bump measuring program. Each line represents<br />

a flat panel light source covered with a thin metal mask with an array of<br />

a shift of a centroid position of each light dot relative<br />

small holes. This assembly was mounted on the downstream crystal holder to its original non-distorted centroid position. The X<br />

and Y components for each dot represent the g<br />

of the monochromator. The CCD camera was mounted on the top of the<br />

x<br />

and g y<br />

components of the gradient vector fields<br />

monochromator box to observe through a viewport the image of the light<br />

dots as reflected off the first monochromator crystal. The CCD images were<br />

captured and analyzed in-situ by a special program based on a modified version of the program Centroid, described earlier in ref.<br />

6. In Figure 1b, the change of the centroid positions of the light dots is shown as a result of the surface distortion. The shift of each<br />

dot characterizes the slope error vector as ∆g= ∆r/L, where ∆r is the shift of a dot’s virtual position and L is the distance between<br />

the virtual image of the light dot and the CCD camera. It is easy to see that the x and y components of the centroid shifts give the<br />

gradient field {g x<br />

, g y<br />

} of the distorted surface.<br />

The reconstruction of the surface from the gradient field has been studied extensively as it is the central issue for a number of<br />

applications; for instance, it is known as the “shape-from-shade” problem. Reconstruction of the shape from the gradient is not a<br />

problem in one dimension but in two dimensions; in this case, the experimentally measured gradient field does not necessarily<br />

satisfy the integrability requirement. An elegant solution of the problem was proposed by A. Agrawal et.al. 7 , by solving the<br />

Poisson equation with the Neumann boundary condition.<br />

An example of the heat-bump formed on the Si(111) crystal, as reconstructed from the measured gradient field, is shown in Figure<br />

2a. In this case, the white x-ray beam from the CHESS A-line wiggler was incident on the monochromator crystal at an angle of<br />

CHESS News Magazine 2009 Page 61


9°. The beam was defined by the 1x6 mm slit (vertical/<br />

horizontal) and the total X-ray power impinging on the Si<br />

crystal was ~ 90 W as measured by a bolometer. In Figure<br />

2b and c, the section of the heat-bump and the slope error<br />

are shown, respectively, along the beam direction through<br />

the heat-bump maximum. The solid lines show the heatbump<br />

and the slope error as derived from the reconstructed<br />

surface, and the crosses are the measured data. The 1D<br />

heat-bump height shown in Figure 2b was obtained by a<br />

simple integration of the experimental slope values. In [8],<br />

this technique has been applied to multilayer optics and<br />

proved that the multilayers on SiC substrates develop a<br />

factor of two smaller heat-bumps than the same multilayers<br />

on traditional Si substrates.<br />

To verify our experimental data we developed a computational algorithm to calculate<br />

the heat-bump. The algorithm is based on the characteristics of the x-ray source,<br />

material characteristics of the first crystal, and the cooling scheme included in the<br />

finite element analysis routine. First, a three-dimensional representation of the x-ray<br />

power absorbed in the Si crystal was calculated by using a Matlab based program.<br />

The program uses the wiggler characteristics and beam-line parameters, including<br />

absorbers and slits, to approximate the 3D power absorption in the crystal by a<br />

number of staggered “bricks” along the beam’s path. Next, a mechanical model of the<br />

Si crystal, plus the copper crystal holder with water cooling, was created in the ANSYS<br />

workbench. Then the 3D heat-load data from the Matlab program were imported<br />

into the ANSYS workbench and the steady-state thermal distribution was calculated<br />

followed by a static structural simulation.<br />

In Figure 3a, the result of the structural simulation is shown. In Figure 3b the<br />

comparison of the heat-bump profile obtained by the ANSYS simulation (black) and<br />

the measured data are shown. Note that the measured data covers only a part of the<br />

Si crystal, determined by the field of view of the optical setup, whereas the ANSYS<br />

simulation generates the profile over the whole Si crystal. The measured and the<br />

simulated results are in a good agreement and they show that the crystal distortion<br />

extends well beyond the beam footprint. We found that the slope error profile obtained<br />

from the ANSYS simulation, and the experimentally measured slope error profile, are<br />

also in a good agreement.<br />

References:<br />

1. D.H. Bilderback, A.K. Freund, G.S. Knapp, and D. M. Mills; “The Historical Development<br />

of Cryogenically Cooled Monochromators for Third-generation <strong>Synchrotron</strong> Radiation<br />

<strong>Source</strong>s“, J. <strong>Synchrotron</strong> Rad. 7, 53-60 (2000)<br />

2. L.E. Berman, J.B. Hastings, D.P. Siddons, M. Koike, V. Stojanoff, and M. Hart;<br />

“Diamond Crystal X-ray Optics for <strong>High</strong>-power-density <strong>Synchrotron</strong> Radiation Beams”,<br />

Nucl. Instr. & Methods A329, 555-563 (1993)<br />

3. A. Kazimirov, D.-M. Smilgies, Q. Shen, X. Xiao, Q. Hao, E. Fontes, D.H. Bilderback, and<br />

S.M. Gruner; “Multilayer X-ray Optics at CHESS “, Journal of <strong>Synchrotron</strong> Radiation,<br />

13, 204-210 (2006)<br />

Fig. 2: The heat-bump surface reconstructed from<br />

the measured gradient vector field using the Poisson<br />

method (a). The heat-bump (b) and the slope error<br />

profile (c) along the beam direction. The solid line<br />

is obtained from the reconstructed surface, and the<br />

red markers represent the measured data.<br />

8. A. Kazimirov, P. Revesz, and R. Huang; “Multilayer Optics under CHESS A2 Wiggler Beam”, Proceedings of SPIE, v. 7077,<br />

707702-1-8 (2008)<br />

Page 62 CHESS News Magazine 2009<br />

(a)<br />

beam footprint<br />

(b)<br />

Fig. 3: (a) ANSYS simulation of heat-bump<br />

and (b) the comparison between the<br />

simulated (black) and the measured (red)<br />

heat-bump profiles for the 90W X-ray power<br />

from the CHESS A2 wiggler impinging at 9°<br />

on the Si monochromator crystal.<br />

4. P. Revesz, A. Kazimirov, and I. Bazarov; “In-situ Visualization of Thermal Distortions of <strong>Synchrotron</strong> Radiation Optics under Wiggler<br />

Beam”, Nucl. Instr. & Methods A576, 422-429 (2007)<br />

5. P. Revesz, A. Kazimirov, and I. Bazarov; “Optical Measurement of Thermal Deformation of Multilayer Optics under <strong>Synchrotron</strong><br />

Radiation“, Nucl. Instr. & Methods A582, 142-145 (2007)<br />

6. P. Revesz, and J. A. White; “An X-ray Beam Position Monitor Based on the Photoluminescence of Helium Gas”, Nucl. Instr. &<br />

Methods A540, 470-479 (2005)<br />

7. A. Agrawal, R. Chellappa, and R. Raskar; “An Algebraic Approach to Surface Reconstruction from Gradient Fields”, IEEE<br />

International Conference on Computer Vision (ICCV), (2005)


Advances in X-ray Microfocusing with<br />

Monocapillary Optics at CHESS<br />

Sterling W. Cornaby 1,2,3 , Thomas Szebenyi 1 , Heung-Soo Lee 4 ,<br />

and Donald H Bilderback 1,2<br />

1<br />

<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>, <strong>Cornell</strong> <strong>University</strong><br />

2<br />

Applied and Engineering Physics, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

Currently at Moxtek Inc., Orem, UT<br />

4<br />

Pohang Accelerator Laboratory, Pohang, Korea<br />

Single-bounce monocapillary optics are ellipticallyshaped<br />

hollow glass tubes which are capable of<br />

focusing x-ray beams at CHESS to a spot size between<br />

5 and 50 µm, with intensity gains ranging from<br />

10 to 1000, and divergences ranging from 1 to 10<br />

milliradians (mr) 1,2,3 . Experiments at CHESS using<br />

x-ray microbeams created with these optics include<br />

high pressure powder diffraction, high resolution<br />

micro-diffraction (µXRD) 4,5 , micro-x-ray fluorescence<br />

(µXRF) 6,7 , micro-XANES 8 , confocal x-ray fluorescence<br />

on antique paintings (confocal µXRF) 9 , micro-protein<br />

crystallography 10 , Laue protein crystallography 11 ,<br />

micro-small angle x-ray scattering (µSAXS) 12 , timeresolved<br />

powder diffraction of reactive multilayer<br />

foils 13 , miniature toroidal mirrors for grazing-incidence<br />

SAXS 14 , micro-X-ray standing waves 18 , and others 15 .<br />

Over the past few years, experiments using<br />

microbeams on various stations at CHESS have<br />

increased greatly. Between 2005 and the present,<br />

CHESS has increased its hardware to accommodate<br />

from one to five simultaneous capillary setups, with<br />

specialty setups for protein crystallography and<br />

μSAXS. The increase in x-ray microbeam capacity<br />

has been fueled by user requests. During a typical<br />

run, one or two x-ray microbeam setups are in use<br />

continually throughout the run, with a peak of four<br />

of CHESS’s twelve stations using x-ray microbeams<br />

simultaneously in November 2007.<br />

Single-bounce monocapillaries are shaped like a small<br />

section of a very eccentric ellipsoid. Rays emitted<br />

from an x-ray source at one focus of an ellipse are<br />

directed to the opposite focus where the sample<br />

under study is placed, as the rays undergo a single<br />

specular reflection from the inner capillary wall. The<br />

ellipsoidal shape is designed to satisfy the grazing<br />

incidence requirement needed for total external<br />

reflection of x-rays. This basic premise allows for<br />

many potential shapes for optics that depend on<br />

the maximum divergence chosen, spot size required<br />

and working length from capillary tip to focus. For<br />

synchrotron applications, the source is typically<br />

located many meters away from the optic and the<br />

incident radiation has a low glancing angle on the<br />

glass wall of order 0.2°.<br />

We evaluate the quality of our drawn optics with 4 kinds of tests.<br />

First, we evaluate the departure of the observed glass figure<br />

from the ideal design shape. This is accomplished using optical<br />

metrology to evaluate the rms slope errors and rms centerline<br />

deviations from a straight line. Second, we observe a far-field<br />

image on a fluorescent screen to measure the divergence of the<br />

capillary and to see if it has the proper ring structure. The farfield<br />

x-ray image is viewed on a high quality fluorescent screen<br />

about 25 to 100 cm downstream from the focus. This far-field<br />

image is used to align the optic with the x-ray beam and contains<br />

information regarding the straightness and the slope errors of<br />

the optic. Third, we take a pinhole scan at the focus of the optic<br />

to see that it makes a small spot. The spot size is measured by<br />

scanning a 5 to 10 μm diameter pinhole across the focus. Fourth<br />

and last, we measure the x-ray beam intensity through the small<br />

pinhole with an ion chamber, with and without a capillary. The<br />

gain is simply the ratio of the two intensity numbers.<br />

Figure 1 shows both a pinhole scan and a far-field image from a<br />

single-bounce monocapillary optic.<br />

Fig 1: A schematic cut away of a single-bounce monocapillary<br />

is in the lower left corner. The positions of the reflected rays are<br />

in dark blue and positions of the non-reflected rays are in light<br />

blue. In the upper left corner is a pinhole scan across the focus<br />

of an optic, with a spot size of 5 μm FWHM, determined after<br />

deconvoluting the data with the 5 μm pinhole size. A far-field<br />

image is on the right side. In the schematic, the far-field screen<br />

is represented by the red line, which is normal to the x-ray<br />

beam. Both the scan and the far-field image are from capillary<br />

f1b_mr9f20_01. Note that the outer ring is not perfectly round.<br />

The strands of concentric intensity arise from several periods of<br />

slight oscillations that develop around the desired figure, most<br />

likely imprinted during the drawing process itself. The diagram<br />

was taken from reference [1].<br />

CHESS News Magazine 2009 Page 63


Single-bounce monocapillaries have a Fig. 2: A diagram of the<br />

number of attributes, listed below.<br />

setup for two total-reflection<br />

mirrors and a single 300 nm<br />

The positive attributes:<br />

thick transmission mirror<br />

• They are achromatic. They reflect all used to create a tunable,<br />

large-bandwidth beam for<br />

x-ray energies to the same focal spot crystallography. The diagram<br />

position.<br />

was redrawn from reference [1].<br />

• They are optically and mechanically<br />

robust. For example, they are not<br />

particularly fragile and they are<br />

typically operated in air.<br />

• They are 90% to 99% efficient. Almost<br />

all the x-rays that hit the surface of the<br />

optic are reflected, if the condition of<br />

total external reflection is satisfied.<br />

• The divergence and focal length can<br />

be designed.<br />

The limiting monocapillary attributes are:<br />

• They have profile and slope errors,<br />

which currently limit the spot size to 5<br />

µm or larger, at CHESS, with the present glass drawing technology.<br />

• They are not imaging optics, they simply focus the beam.<br />

• They have finite aperture size of about 1 mm or less and divergence<br />

of about 12 mrad or less. The small numerical aperture limits the<br />

amount of x-ray light that the optic can collect.<br />

Being achromatic, however, is an exceptionally important advantage of<br />

the monocapillary optics. This means the focal spot size, the focal spot<br />

location, and the divergence of the focused beam do not change with<br />

the x-ray photon energy. Thus these optics focus a 0.01% bandwidth<br />

beam as effectively as a 50% bandwidth beam. The reflection efficiency<br />

remains high over the entire x-ray energy range as long as the x-rays do<br />

not exceed the critical angle.<br />

A recent example of using this achromatic principle to good advantage can be<br />

seen in a Laue diffraction experiment whose purpose is to determine the 3D<br />

crystal structure of protein microcrystals 1,16 . This experiment used a combination<br />

Page 64 CHESS News Magazine 2009<br />

Fig. 3: A graph showing the predicted<br />

efficiency and measured spectrum of the<br />

30% bandwidth (FWHM) created from two<br />

rhodium reflection mirrors and one Si 3<br />

N 4<br />

transmission mirror combination. The pink<br />

curve is calculated, and the blue-green curve<br />

is measured. They compare favorably. The<br />

bandwidth is peaked at 12 keV (λ=1.24 Å)<br />

with half-height values at 9.6 keV (λ=1.29 Å)<br />

and 13.4 keV (λ=0.93 Å). In the measured<br />

Compton scattering curve (green), the several<br />

sharp fluorescence peaks originate from trace<br />

contaminants within the Kapton® tape; they<br />

are not present in the incident X-ray beam.<br />

The diagram was taken from reference [1].<br />

Fig. 4: Right, Laue diffraction pattern from a<br />

lysozyme crystal. Left, magnified view of the area<br />

in the magenta box, showing well-separated,<br />

acceptably-shaped spots.<br />

of two rhodium coated reflection mirrors and a silicon nitride x-ray transmission mirror to generate a 30% bandwidth beam<br />

peaked at 12 keV, Figures 2 and 3.<br />

The wide-bandwidth beam was then focused with a single-bounce monocapillary optic to a 10(H)x13(V) µm 2 spot, with a<br />

divergence of ~ 5(H)x2(V) mrad 2 and resulted in a total flux of 4.4×10 10 photons/s (~3.4×10 8 photons/s/µm 2 over the 30%<br />

bandwidth). This experiment was performed at the D1 bending magnet station whose critical energy is about 10 keV.<br />

The flux density achieved using the bending magnet is close to what you would expect for a 1-2% bandwidth microfocused beam<br />

on a wiggler station. This small beam was then used to take Laue diffraction images of small protein crystals, mostly lysozyme, but<br />

also of a few thaumatin crystals, Figure 4.<br />

The crystals were a few tens of microns in size. In the short term, we believe that this approach will lead to a new avenue for<br />

solving protein structures using less than 10 µm sized crystals at CHESS. In the longer term at an ERL source of x-rays, we hope to<br />

push crystal sizes down into the hundreds of nanometers range with this approach - an approach that will be more limited by the<br />

radiation damage of crystals than by any beam qualities.<br />

Another important advantage of single-bounce monocapillary optics is that the divergence and focal length can be designed.<br />

CHESS has a glass capillary puller which has been set up to make a number of different optics 1 . Two examples show the power<br />

of having this in-house tool for constructing specialized monocapillary optics. We have been able to fabricate “football” shaped<br />

optics with very short focus-to-focus distances of 20 to 30 cm 1 . We have also been able to make a bifocal miniature toroidal mirror<br />

that horizontally and vertically focuses to two different locations 14 .<br />

The need for a football monocapillary optic arose within the Confocal X-Ray Fluorescence (CXRF) project, led by CHESS staff<br />

scientist Arthur Woll and his collaborators, to investigate buried paint layer structures of antique paintings 9 . CXRF uses x-ray


optics to define a small viewable volume<br />

in free space that the XRF detector can see.<br />

By limiting the volume of the detection<br />

region with two separate x-ray optics, the<br />

third dimension (the depth) of the sample’s<br />

elemental composition can be resolved,<br />

such as in a painting’s underlying layers. At<br />

CHESS, a single-bounce monocapillary optic<br />

is used to focus the x-ray beam and an XOS<br />

polycapillary optic is used to collect XRF<br />

events for the detector, Figure 5.<br />

We decided to try CXRF with two crossed<br />

monocapillary optics. The second collection<br />

optic needed to have a very short focus-tofocus<br />

distance of 30 cm or less. We<br />

made this second football-shaped<br />

optic and were able to use it in a<br />

CXRF experimental setup to test<br />

the confocal resolution, Figure 6.<br />

The football-shaped optic is a<br />

good fit for x-ray tube sources,<br />

where short focus-to-focus<br />

distances improve the total usable<br />

x-ray flux from the tube. We are<br />

presently looking further into other<br />

applications using this kind of<br />

optic.<br />

We have also made a non-elliptically shaped<br />

single-bounce monocapillary optic at CHESS.<br />

A special dual-focal length toroidal mirror<br />

Fig. 5: The left diagram shows how a small detection volume in space is created at the intersection<br />

of a beam produced by the monocapillary and accepted by the polycapillary optic. The confocal<br />

volume is green in color in the inset. The diagram on the right shows how this small volume can<br />

be used to resolve a layered structure. Since the confocal volume is about the same dimension as a<br />

typical paint layer on an oil painting, the composition of the different layers can be resolved in three<br />

dimensions. The diagram was taken from reference [9].<br />

Fig. 6: The left diagram shows<br />

a sketch of the CXRF test<br />

made with two monocapillary<br />

optics. The confocal volume<br />

was measured with an XRF<br />

detector, as a 6 µm thin lead<br />

foil was passed through<br />

the confocal volume. Spot<br />

defining slits at the second<br />

focus of the football shaped<br />

collection optic controlled the<br />

viewable size of the confocal<br />

volume. The graph on the<br />

right shows the resolution as a function of the slit opening. A resolution of ~120 µm FWHM was<br />

achieved with a 300 µm slit setting, and a resolution of ~50 µm FWHM was achieved with a 50 µm<br />

slit setting. The diagram was taken from reference [1].<br />

was designed to focus vertically at the sample’s position 50 mm from the capillary tip and to horizontally focus at the detector’s<br />

position, 150 mm from the capillary tip. This mirror was made to provide a smaller vertical footprint of beam for grazing incidence<br />

wide-angle scattering (GIWAXS), while at the same time producting better angular resolution in the horizontal direction. No other<br />

capillary has been designed to date with such separate focal distances!<br />

This optic, however, is very similar in concept to the elliptically-shaped single-bounce monocapillary optics. This miniature<br />

toroidal mirror was designed to decouple the sagittal and meridional focusing of the traditional single-bounce monocapillary<br />

optic, Figure 7.<br />

Essentially, the normal ellipsoidal shape is modified by increasing its inner diameter, changing the sagittal focus, and at the same<br />

time not modifying its meridional curvature, leaving the meridional focusing capability unchanged.<br />

Using the entire inner optical surface will not produce the needed horizontal line focus at the sample position and a vertical line<br />

focus at the detector position. In order to create a line focus, only a small section of the optic’s full inner surface is exposed to<br />

x-rays. Slits are set upstream of the optic to block most of the optical surface illumination as only about 10% of the inner optical<br />

surface area is used. [If the full surface of the optic is exposed to an x-ray beam, a narrow ring of intensity will be observed at the<br />

location of the meridional focus, and a<br />

semi-large spot will appear at the sagittal<br />

focused position.] X-ray tests were<br />

made with this optic, Figure 8.<br />

Improvements have been made to the<br />

capillary drawing process. Parts are now<br />

drawn under constant pressure (rather<br />

than constant tension) and the furnace<br />

moves in the opposite direction to the<br />

upward drawing of glass out of the<br />

heating zone, Figure 9.<br />

These changes, instituted by Heung-Soo<br />

Lee when he was a visiting scientist with<br />

Fig. 7: This sketch outlines the morphing of a ellipsoidal shape (left) into a toroidal shape (right). The<br />

diameter of the optic is changed, thereby decoupling, the sagittal and meridional focal points. The<br />

meridional focus position is unchanged, but the sagittal focusing is moved further away from the tip of<br />

the optic. The diagram was taken from reference [14].<br />

CHESS News Magazine 2009 Page 65


us last year, have contributed<br />

to drawing better optics.<br />

In addition, H-S. Lee also<br />

worked on making far-field<br />

simulations from the slope<br />

errors measured from the<br />

on-board Keyence optical<br />

metrology, Figure 10.<br />

In conclusion, micro-focusing<br />

single-bounce monocapillary<br />

optics will continue to be an integral part of the capabilities of CHESS.<br />

Work will continue to perfect the monocapillary optics by reducing<br />

slope and figure errors in the fabrication process. With improvements<br />

in slope and figure errors, these optics should make smaller beam<br />

sizes and will be very useful for both 3rd generation x-ray sources and<br />

the ERL.<br />

Fig. 8: The beam was observed (with 5 micron<br />

diameter pinhole scans) to be horizontally focused<br />

150 mm from the capillary tip and vertically focused<br />

only 50 mm from the tip, proof that the dual-focal<br />

lengths had been produced. Some structure can<br />

be seen in the vertical scan, showing that more<br />

perfecting work is needed to turn this prototype<br />

into a usable optic. Nonetheless, this is an excellent<br />

demonstration that a new type of capillary optic<br />

has been created! The diagram was taken from<br />

reference [14].<br />

Fig. 9: Capillary drawing tower<br />

showing AB Tech air-bearing<br />

in the background, tensionproducing<br />

stage on the left,<br />

furnace and optical metrology<br />

in the foreground (with<br />

suspended glass tube) and a<br />

load cell at the bottom of the<br />

apparatus.<br />

Fig. 10: Far-field capillary image (including direct beam in center).<br />

Circular strands of intensity are due to small slope errors remaining<br />

from the glass drawing process. The simulation on the right shows<br />

similar ring structure and is approaching the circular structures seen<br />

at left when realistic slope-error values are included in the calculation.<br />

Acknowledgement:<br />

We wish to acknowledge that our work was supported by CHESS<br />

through the NSF & NIH/NIGMS via NSF award DMR-0225180 and by the<br />

MacCHESS resource through NIH/NCRR award RR-01646. One of us, SC,<br />

was also supported by a <strong>Cornell</strong> <strong>University</strong> supported G-line fellowship.<br />

References:<br />

1. S. Cornaby; The Handbook of X-ray Single-Bounce Monocapillary<br />

Optics, Including Optical Design and <strong>Synchrotron</strong> Applications,<br />

Dissertation, <strong>Cornell</strong> <strong>University</strong> 2008. Available at http://glasscalc.<br />

chess.cornell.edu/Cornaby_Capillary_Handbook_2008.pdf.<br />

2. S. Cornaby, T. Szebenyi, R. Huang, and D.H. Bilderback; “Design<br />

of Single-Bounce Monocapillary X-Ray Optics”, 55th Denver X-ray<br />

Conference, Advances in X-ray Analysis Vol. 50 (2006)<br />

3. R. Huang and D. H. Bilderback; “Single-bounce Monocapillaries<br />

for Focusing <strong>Synchrotron</strong> Radiation: modeling, measurements and<br />

theoretical limits”, J. of <strong>Synchrotron</strong> Radiation 13, 74-84 (2006)<br />

4. A.A. Sirenko, A. Kazimirov, S. Cornaby, D.H. Bilderback, B. Neubert,<br />

P. Brueckner, F. Scholz, V. Shneidman, and A. Ougazzaden;<br />

“Microbeam <strong>High</strong> Angular Resolution X-ray Diffraction in InGaN/GaN<br />

Selective-area-grown Ridge Structures”, Applied Physics Letters 89<br />

(18): Art. No. 181926 (Oct. 30 2006)<br />

5. A. Kazimirov, A.A. Sirenko, D.H. Bilderback, Z.-H. Cai, and B.<br />

Lai; “Microbeam <strong>High</strong> Angular Resolution Diffraction Applied to<br />

Optoelectronic Devices”, AIP Conference Proceeding CP879,<br />

<strong>Synchrotron</strong> Radiation Instrumentation: Ninth International<br />

Conference 1395-1397 (2007)<br />

6. K. Limburg, R. Huang, and D.H. Bilderback; “Fish Otolith Trace<br />

Element Maps: new approaches with synchrotron microbeam X-ray<br />

fluorescence”, X-ray Spectrometry, 36, 336-342 (2007)<br />

7. C. Schmidt, K. Rickers, D.H. Bilderback, and R. Huang; “In situ<br />

<strong>Synchrotron</strong>-radiation XRF Study of REE Phosphate Dissolution in<br />

Aqueous Fluids to 800°C”, Lithos 95, 87-102 (2007)<br />

Page 66 CHESS News Magazine 2009<br />

8. R.A. Barrea, R. Huang, S. Cornaby, D.H. Bilderback, and T.C.<br />

Irving; “<strong>High</strong>-flux Hard X-ray Microbeam using a Single-bounce<br />

Capillary with Doubly Focused Undulator Beam”, J. of <strong>Synchrotron</strong><br />

Radiation 16, 76-82 (2009)<br />

9. A.R. Woll, J. Mass, C. Bisulca, R. Huang, D.H. Bilderback, S. Gruner,<br />

and N. Gao; “Development of Confocal X-ray Fluorescence (XRF)<br />

Microscopy at the <strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>”, Applied<br />

Physics A 83 (2), 235-238 (2006)<br />

10. R. Huang and D.H. Bilderback; “Secondary Focusing for Micro-<br />

Diffraction using One-Bounce Capillaries”, Eighth International<br />

Conference on <strong>Synchrotron</strong> Radiation Instrumentation, AIP<br />

Conference Proceedings, 712-715 (2004)<br />

11. Ibid ref. 1<br />

12. J.S. Lamb, S. Cornaby, K. Andresen, L. Kwok , H.Y. Park, X. Qiu,<br />

D.M. Smilgies, D.H. Bilderback, and L. Pollack; “Focusing Capillary<br />

Optics for use in SAXS”, Journal of Applied Crystallography 40,<br />

193-195 (2007)<br />

13. J.C. Trenkle, L.J. Koerner, M.W. Tate, S.M. Gruner, T.P. Weihs, and<br />

T.C. Hufnagel; “Phase Transformations during Rapidly Propagating<br />

Reactions in Nanolaminate Foils”, App. Phy. Lett. vol. 92 (2008)<br />

14. S. Cornaby, D. Smilgies, and D. Bilderback; “Bifocal Miniature<br />

Toroidal Shaped X-ray Mirrors”, 57th Denver X-ray Conference,<br />

Advances in X-ray Analysis Vol. 52 (in print, 2009)<br />

15. D.H. Bilderback, A. Kazimirov, R. Gillilan, S. Cornaby, A. Woll,<br />

C-S Zha, and R. Huang; “Optimizing Monocapillary Optics<br />

for <strong>Synchrotron</strong> X-ray Diffraction, Fluorescence Imaging, and<br />

Spectroscopy Applications”, AIP Conference Proceeding CP879,<br />

<strong>Synchrotron</strong> Radiation Instrumentation: Ninth International<br />

Conference, 758-763 (2007)<br />

16. S. Cornaby, D.M.E. Szebenyi, D-M. Smilgies, D.J. Schuller, R.<br />

Gillilan, Q. Hao, and D.H. Bilderback, Acta Cryst. D. (submitted)<br />

17. S. Cornaby and D.H. Bilderback; “Silicon Nitride Transmission X-ray<br />

Mirrors”, Journal of <strong>Synchrotron</strong> Radiation 15, 371-373 (2008)<br />

18. A. Kazimirov, D.H. Bilderback, R. Huang, A. Sirenko, and A.<br />

Ougazzaden; “Microbeam <strong>High</strong>-resolution Diffraction and X-ray<br />

Standing Wave Methods Applied to Semiconductor Structures”, J.<br />

Phys. D: Appl. Phys. 37, 3L9-L12 (2004)


Watching Carbon Nanotube Forest Growth<br />

using X-rays<br />

Eric R. Meshot 1 , Mostafa Bedewy 1 , Sameh Tawfick 1 , K. Anne Juggernauth 1 ,<br />

Eric Verploegen 2 , Yongyi Zhang 1 , Michael De Volder 1 , and A. John Hart 1<br />

1<br />

Department of Mechanical Engineering, <strong>University</strong> of Michigan<br />

2<br />

Department of Materials Science and Engineering, MIT<br />

Carbon nanotubes (CNTs) have many<br />

potential applications due to their atomic<br />

structure and consequent mechanical 1 ,<br />

electrical 2 and thermal 3 properties.<br />

CNTs can be classified into two distinct<br />

structural forms: (i) a single seamless<br />

cylindrical shell of sp 2 -bonded carbon<br />

atoms arranged in a honeycomb lattice<br />

constituting a single-wall CNT (SWCNT)<br />

which typically has a diameter ranging<br />

from 0.4 nm to 3nm; (ii) several concentric<br />

seamless shells with 0.34 nm interwall<br />

spacing, forming a multi-wall CNT<br />

(MWCNT) which can have diameters<br />

ranging from 1.4 nm to 100 nm. An<br />

infinitely long assembly of aligned,<br />

continuous, and densely packed CNTs<br />

would constitute the ultimate of synthetic<br />

fibers, and could have several times the<br />

strength of piano wires at one-fourth<br />

the density, along with thermal and<br />

electrical conductivity exceeding copper.<br />

However, before this dream is realized,<br />

many smaller-scale configurations of CNTs<br />

are potentially useful for next-generation<br />

transistors, flexible and transparent<br />

electronics, and multifunctional interface<br />

layers. Specifically, vertically aligned CNT<br />

“forests”, consisting of tens of billions<br />

of CNTs per square centimeter on a flat<br />

substrate, are widely sought as highperformance<br />

electrical, thermal, and<br />

mechanical interface layers, as well as<br />

filtration membranes 4-10 .<br />

Although CNTs can be synthesized<br />

by many methods involving hightemperature<br />

decomposition and<br />

reorganization of solid carbon, the most<br />

popular and scalable means of CNT<br />

synthesis is catalytic chemical vapor<br />

deposition, where CNTs “grow” from metal<br />

catalyst nanoparticles that are typically<br />

placed on a substrate or floated in a<br />

carbon-containing gas atmosphere. The<br />

process of CNT forest growth by thermal<br />

CVD typically involves multiple stages: (1)<br />

the catalyst is prepared on a substrate,<br />

such as a silicon wafer; (2) the catalyst is heated and treated chemically, such as by<br />

exposure to a reducing atmosphere that causes agglomeration of a thin film into<br />

nanoparticles; (3) the catalyst is exposed to a carbon-containing atmosphere, which<br />

causes formation and “liftoff” of CNTs from the nanoparticles on the substrate; and<br />

(4) CNT growth continues by competing pathways between accumulation of “good”<br />

(graphitic) and “bad” (amorphous) carbon 11 . To engineer the functional properties<br />

of CNT materials such as forests, we must develop reaction processes that not only<br />

treat these stages independently, but that are also accompanied by characterization<br />

techniques that enable mapping of the forest characteristics for large sample sizes<br />

and populations. Despite extensive study of aligned CNT forest production 12-27 ,<br />

including recent advances using water and oxygen as additives to increase reaction<br />

yield and catalyst lifetime, the limiting mechanisms of forest growth are not fully<br />

understood, and CNT forest heights are typically limited to several millimeters.<br />

In Figure 1, we show an AFM image of nanoparticles arranged for CNT forest<br />

growth, as well as images of a vertically aligned multi-wall CNT forest at different<br />

magnifications.<br />

Fig. 1: (a) 500 nm AFM scan of Fe nanoparticles formed on an<br />

Al 2<br />

O 3<br />

support layer by heating in H 2<br />

for 2 minutes at 825 °C; and<br />

vertically aligned multi-wall CNTs at different magnifications: (b) 10 X,<br />

(c) 140,000 X (SEM) and (d) 7,000,000 X (TEM).<br />

At CHESS, we use ex situ and in situ small-angle X-ray scattering (SAXS) to investigate<br />

growth of CNT forests, and work closely with Dr. Arthur Woll, Prof. Sol Gruner, and<br />

colleagues. In order to elucidate the dynamics of catalyst particle coarsening, forest<br />

self-organization, temperature- and reactant-driven CNT structure evolution, and<br />

growth termination in realtime, a custom-built atmospheric-pressure CVD reactor<br />

featuring a resistively heated substrate platform 28 , shown in Figure 2, is mounted<br />

directly in the G1 beamline at CHESS, enabling in situ grazing incidence (GI-SAXS)<br />

and transmission (SAXS, WAXS) scattering studies, as shown in Figure 3. Simultaneous<br />

laser measurement of the forest height captures the growth kinetics, the<br />

heated platform enables rapid temperature changes (200 °C/s) during annealing<br />

and growth, and the reactant gas is independently heated to create a population of<br />

active carbon species that cause efficient CNT growth. Further, Figure 3 summarizes<br />

the different techniques and resulting data for both GI and transmission modes: we<br />

show schematics of our setups, examples of scattering intensities collected by the<br />

FLICAM detector, and data demonstrating temperature effects on nanoparticle and<br />

CNT formation.<br />

CHESS News Magazine 2009 Page 67


Fig. 2: Photograph of CNT<br />

growth apparatus mounted<br />

in the G1 Beamline.<br />

Fig. 3: Schematic and data<br />

for both grazing incidence<br />

(a-c) and transmission (d-f)<br />

SAXS configurations. For<br />

GI mode: (a) schematic<br />

shows X-rays glancing off<br />

substrate-bound particles;<br />

(b) scattering intensity<br />

from a population of Fe<br />

nanoparticles forming in<br />

real time as temperature<br />

increases; (c) I vs. q profiles<br />

showing real-time evolution<br />

of the structure factor peak<br />

for the Fe nanoparticles.<br />

For transmission mode: (d) schematic shows X-rays passed directly through a CNT forest; (e)<br />

representative scattering intensity from a CNT forest; (f) I vs. q profiles with curve fits showing the<br />

dependence of the CNT form factor peak on growth temperature – the shift of the peak position<br />

to lower q indicates larger CNT mean diameter.<br />

After growth, we also carry out ex situ characterization of the CNT forests in order to spatially map their bulk morphology 29 . Using<br />

SAXS we can investigate the spatial variations in MWCNT orientation. In addition we use SAXS to measure a locally averaged spatial<br />

variation in CNT diameters within our films. CNT diameters obtained by fitting the scattering results were confirmed by TEM<br />

imaging, which establishes SAXS as an attractive non-destructive technique for precisely examining CNT materials 29,30 .<br />

Starting with a catalyst thin-film of Fe/Al 2<br />

O 3<br />

on Si, we observe rapid coarsening of Fe at temperatures as low as 650 °C (Figure<br />

3c) 31 . CNT diameter and growth rate are directly proportional to the substrate temperature (Figure 3f) 30 , and tuning of annealing<br />

and growth conditions using SAXS-derived diameter measurements reveals that CNT forests with mean diameters ranging from<br />

4-20 nm can be grown from the same starting catalyst film thickness (not shown here) 31 . Growth self-terminates abruptly, accompanied<br />

by a sudden loss of alignment at the CNT-substrate interface (Figure 4); this appears to be a universal chemical and/or<br />

mechanical signature in our experiments 32 . Our apparatus and investigative technique offer significant potential to further understand<br />

the limiting mechanisms of CNT forest growth, and for rapid tuning of process conditions to engineer application-oriented<br />

structural characteristics of nanotubes and nanowires.<br />

Figure 4a shows in situ measurements of the height evolution of a forest that self-terminates, and of forests which are terminated<br />

prematurely by rapidly cooling the substrate or stopping the flow of the carbon containing gas while maintaining the substrate<br />

temperature 33 . As seen in Figure 4b, the orientation (quantified by the Hermans orientation parameter) increases sharply at the<br />

top of a forest, representing the transition from tangled to vertically aligned morphology during the first stage of growth. The<br />

orientation then remains approximately constant as growth proceeds, and then it decays steeply toward zero before growth<br />

terminates, indicating the onset of disordered CNTs. In agreement with the morphological evolution indicated by the calculated<br />

Hermans values, SEM images (Figure 4c) show that the self-terminated forest exhibits disorder at its base, yet the intentionally<br />

terminated forests exhibit strong alignment at the base. This demonstrates that the loss of alignment is a signature of growth<br />

self-termination, and is not caused by cooling the reactor or by an abrupt decrease in the carbon concentration in the CVD atmosphere.<br />

Page 68 CHESS News Magazine 2009


Based on our in situ and ex situ results, we explain the evolution and termination<br />

of CNT forest growth based on a collective model consisting of four stages 33 , as<br />

illustrated in Figure 5:<br />

I. nucleation and self-assembly of randomly oriented CNTs into a<br />

vertically aligned forest structure;<br />

II. steady growth, wherein the number density of growing CNTs<br />

remains constant with time;<br />

III. density decay, wherein the number density of CNTs within the forest<br />

decreases with time; and<br />

IV. abrupt termination, wherein forest growth ceases suddenly, and is<br />

accompanied by a loss of CNT alignment at the interface between<br />

the CNTs and the substrate.<br />

Fig. 4: (a) Growth kinetics for 3 CNT forests<br />

for different growth times under otherwise<br />

identical conditions are plotted along with<br />

theoretical quadratic and exponential decay<br />

curves. (b) Relative alignment within each<br />

forest is quantified, and (c) corresponding<br />

SEM images verify the CNT morphology.<br />

In conclusion, we have employed nondestructive X-ray techniques for rapid<br />

characterization of CNT forests, revealing dynamics of their growth and termination<br />

processes and enhancing our ability to better engineer the characteristics<br />

of these forests (i.e., length, density, CNT diameter). Further, we have developed<br />

in situ GI methods with our custom CVD apparatus for probing the dynamics<br />

of catalyst particle formation from thin films for CVD of CNTs. By observing the<br />

agglomeration and coarsening of these particles in real time, we elucidate this<br />

rapid process as we move toward understanding the different chemical states<br />

and transitions of the particles as well as the prerequisites for stabilizing populations<br />

of small particles for small-diameter CNT growth. These applied techniques<br />

may also be adapted for comprehensive studies of other systems of 1-dimensional<br />

(nanotubes, nanowires) and 0-dimensional (nanoparticles) structures,<br />

demonstrating the impact and versatility of such characterization methods for<br />

understanding how to create new materials and nanostructures.<br />

The Mechanosynthesis Group’s X-ray team (Figure 6) makes the following acknowledgements.<br />

This work was funded by the <strong>University</strong> of Michigan Department<br />

of Mechanical Engineering and College of Engineering, and the National<br />

Science Foundation (CMMI-0800213). E. R. Meshot and S. Tawfick are grateful<br />

for <strong>University</strong> of Michigan Mechanical Engineering Departmental Fellowships.<br />

E. A. Verploegen is grateful to the Institute for<br />

Soldier Nanotechnologies at MIT, funded by the U.S.<br />

Army Research Office (DAAD-19-02-D0002). Some<br />

studies (Figure 3f) based from methods developed<br />

at CHESS were conducted at the National <strong>Synchrotron</strong><br />

Light <strong>Source</strong>, Brookhaven National Laboratory,<br />

which is supported by the U.S. Department of<br />

<strong>Energy</strong> Office of Basic <strong>Energy</strong> Sciences (DE-AC02-<br />

98CH10886). Otherwise, X-ray scattering was performed<br />

at CHESS, which is supported by the National<br />

Science Foundation and the National Institutes<br />

of Health under Grant No. DMR-0225180. SEM and<br />

TEM were performed at the <strong>University</strong> of Michigan<br />

Electron Microbeam Analysis Laboratory (EMAL).<br />

We thank Jong G. Ok, Sangwoo Han, Myounggu<br />

Park, Arthur Woll, Mark Tate, Hugh Philipp, Marianne<br />

Hromalik, and Sol Gruner for assistance with X-ray<br />

scattering experiments and analysis.<br />

Fig. 5: Collective growth mechanism of a CNT forest:<br />

(a) growth stages; and SEM images of the tangled<br />

crust at top of forest, aligned morphology at bottom<br />

during steady growth, and the randomly oriented<br />

morphology at bottom, induced by density decay<br />

and abrupt self-termination; (b) time evolution of<br />

Hermans orientation parameter for forests grown for<br />

different durations in a “hot wall” tube furnace..<br />

CHESS News Magazine 2009 Page 69


Fig. 6: Mechanosynthesis Group’s X-ray team in the<br />

hutch at G1 line.<br />

From left to right: Mostafa Bedewy, Michael de Volder,<br />

Eric Verploegen, John Hart, Penguin, Yongyi Zhang,<br />

Eric Meshot, and Sameh Tawfick.<br />

References:<br />

1. N. Yao, and V. Lordi; Journal of Applied Physics 1998, 84 ( 4),<br />

1939-1943 (1998)<br />

2. H.J. Li, W.G. Lu, J.J. Li, X.D. Bai, and C.Z. Gu; Physical Review<br />

Letters, 95 (8), 086601 (2005)<br />

3. P. Kim, L. Shi, A. Majumdar, and P.L. McEuen; Physical Review<br />

Letters, 87 (21), 215502 (2001)<br />

4. R.H. Baughman, A.A. Zakhidov, and W.A. de Heer; Science, 297 ( 5582), 787-792 (2002)<br />

5. M. Endo, T. Hayashi, Y. Kim, M. Terrones, and M. Dresselhaus; Philosophical Transactions of the Royal Society of London Series<br />

A-Mathematical Physical and Engineering Sciences, 362 (1823), 2223-2238 (2004)<br />

6. E.J. Garcia, B.L. Wardle, and A.J. Hart; Composites Part a-Applied Science and Manufacturing, 39 ( 6), 1065-1070 (2008)<br />

7. T. Tong, Y. Zhao, L. Delzeit, A. Kashani, M. Meyyappan, and A. Majumdar; Transactions on Components and Packaging Technologies, 30 ( 1),<br />

92-100 (2007)<br />

8. W. Choi, D. Chung, J. Kang, H. Kim, Y. Jin, I. Han, Y. Lee, J. Jung, N. Lee, G. Park, and J. Kim; Applied Physics Letters, 75 ( 20), 3129-3131(1999)<br />

9. J. Holt, H. Park, Y. Wang, M. Stadermann, A. Artyukhin, C. Grigoropoulos, A. Noy, and O. Bakajin, Science, 312 ( 5776), 1034-1037 (2006)<br />

10. B. Hinds, N. Chopra, T. Rantell, R. Andrews, V. Gavalas, and L. Bachas; Science, 303 ( 5654), 62-65 (2004)<br />

11. A.I. Lacava, C.A. Bernardo, and D.L. Trimm, Carbon, 20 ( 3), 219-223 (1982)<br />

12. L. Dell'Acqua-Bellavitis, J. Ballard, P. Ajayan, and R. Siegel; Nano Letters, 4 ( 9), 1613-1620 (2004)<br />

13. E. Einarsson, Y. Murakami, M. Kadowaki, and S. Maruyama; Carbon, 46 ( 6), 923-930 (2008)<br />

14. D. Futaba, K. Hata, T. Yamada, K. Mizuno, M. Yumura, and S. Iijima; Physical Review Letters, 95 ( 5) (2005)<br />

15. J. Han, R. Graff, B. Welch, C. Marsh, R. Franks, and M. Strano; ACS Nano, 2 ( 1), 53-60 (2008)<br />

16. K. Liu, K. Jiang, C. Feng, Z. Chen, and S. Fan; Carbon, 43 ( 14), 2850-2856 (2005)<br />

17. O. Louchev, T. Laude, Y. Sato, and H. Kanda; Journal of Chemical Physics; 118 ( 16), 7622-7634 (2003)<br />

18. A. Puretzky, D. Geohegan, S. Jesse, I. Ivanov, and G. Eres; Applied Physics A-Materials Science & Processing, 81 ( 2), 223-240 (2005)<br />

19. R. Xiang, Z. Yang, Q. Zhang, G. Luo, W. Qian, F. Wei, M. Kadowaki, E. Einarsson, and S. Maruyama; Journal of Physical Chemistry C, 112 ( 13),<br />

4892-4896 (2008)<br />

20. L. Zhu, J. Xu, F. Xiao, H. Jiang, D. Hess, and C. Wong; Carbon, 45 ( 2), 344-348 (2007)<br />

21. L. Zhu, D. Hess, and C. Wong; Journal of Physical Chemistry B, 110 ( 11), 5445-5449 (2006)<br />

22. G.Y. Zhang, D. Mann, L. Zhang, A. Javey, Y.M. Li, E. Yenilmez, Q. Wang, J.P. McVittie, Y. Nishi, J. Gibbons, and H.J. Dai; Proceedings of the<br />

National Academy of Sciences of the United States of America, 102 (45), 16141-16145 (2005)<br />

23. K. Hata, D.N. Futaba, K. Mizuno, T. Namai, M. Yumura, S. Iijima; Science, 306 ( 5700), 1362-1364 (2004)<br />

24. S.S. Fan, M.G. Chapline, N.R. Franklin, T.W. Tombler, A.M. Cassell, and H.J. Dai; Science, 283 ( 5401), 512-514 (1999)<br />

25. L. Delzeit, C.V. Nguyen, B. Chen, R. Stevens, A. Cassell, J. Han, and M. Meyyappan; Journal of Physical Chemistry B, 106 ( 22), 5629-5635 (2002)<br />

26. A. Hart, and A. Slocum; Journal of Physical Chemistry, 110 ( 16), 8250-8257 (2006)<br />

27. R.F. Wood, S. Pannala, J.C. Wells, A.A. Puretzky, and D.B. Geohegan; Physical Review B, 75 ( 23), 8 (2007)<br />

28. L. van Laake, A.J. Hart, and H.H. Slocum; Review of Scientific Instruments, 78 ( 8) (2007)<br />

29. B.N. Wang, R.D. Bennett, E. Verploegen, A.J. Hart, and R.E. Cohen; Journal of Physical Chemistry C, 111 ( 16), 5859-5865 (2007)<br />

30. E.R. Meshot, D.L. Plata, S. Tawfick, E.A. Verploegen, and A.J. Hart; (Submitted 2009)<br />

31. A.J. Hart, E. Verploegen, and E.R. Meshot; (In preparation 2009)<br />

32. E. Meshot, and A. Hart; Applied Physics Letters, 92 ( 11) (2008)<br />

33. M. Bedewy, E.R. Meshot, H. Guo, E.A. Verploegen W. Lu, and A.J. Hart; (Submitted 2009)<br />

Page 70 CHESS News Magazine 2009


Phase Behavior of Water inside Protein Crystals<br />

Chae Un Kim 1 , Buz Barstow 2 , Mark W. Tate 3 , and Sol M. Gruner 1,3,4<br />

1<br />

Macromolecular Diffraction Division of <strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong>,<br />

<strong>Cornell</strong> <strong>University</strong><br />

2<br />

School of Applied Physics, <strong>Cornell</strong> <strong>University</strong><br />

3<br />

Laboratory of Atomic and Solid State Physics, <strong>Cornell</strong> <strong>University</strong><br />

4<br />

Physics Department, <strong>Cornell</strong> <strong>University</strong><br />

Protein crystals typically consist of ≈ 40 - 60 % water. The internal water forms solvent channels (≈ 2 - 4 nm in<br />

diameter) inside the crystals. We have studied the phase behavior of water inside protein crystals using a novel<br />

crystal freezing method: high-pressure cryocooling. Using x-ray diffraction, we demonstrated that the high-density<br />

amorphous (HDA) ice induced inside the high-pressure cryocooled protein crystal undergoes a phase transition to<br />

low-density amorphous (LDA) ice as the crystal is warmed from 80 to 170 K. We also found evidence for a liquid state<br />

of water during the ice transition. These results have significant implications for the understanding of the anomalous<br />

behavior of supercooled water.<br />

Protein crystals used in x-ray protein<br />

crystallography have properties that<br />

distinguish them from typical inorganic<br />

crystals. An important difference is that protein<br />

crystals contain internal water, typically<br />

amounting to 40 to 60 % of the volume of<br />

the crystal. This internal water forms solvent<br />

channels inside protein crystals, as shown in<br />

Figure 1, and fills the significant void volume<br />

between the folded protein molecules that<br />

make up the crystal.<br />

We have recently studied the phase behavior<br />

of water confined inside protein crystals 1<br />

using a crystal freezing method, high-pressure<br />

cryocooling 2 , which was originally developed<br />

for protein cryoprotection and diffraction<br />

phasing 3,4 . Our previous study revealed that<br />

the high-density amorphous (HDA) ice induced<br />

by high-pressure cryocooling is responsible for<br />

the high quality diffraction that is frequently<br />

obtainable from high-pressure cryocooled<br />

crystals 5 .<br />

Figure 2 shows the diffraction images of<br />

a high-pressure cryocooled crystal of the<br />

globular protein thaumatin, as the crystal is<br />

warmed from 80 to 170 K. By filtering out the<br />

crystal Bragg diffraction from these images,<br />

the water diffuse diffraction (WDD) was<br />

isolated, allowing the phase behavior of water<br />

present in the internal solvent channels to be<br />

studied. At the same time, the crystal response<br />

(variation of unit-cell parameters and crystal<br />

mosaicity, etc) and the protein molecular<br />

response to the phase transition of the solvent<br />

channel water could be studied by analysis of<br />

the crystal Bragg diffraction.<br />

Fig. 1: Solvent channels of a thaumatin crystal oriented along the crystallographic<br />

a axis. The channel diameter is approximately 3 nm. The solvent inside the channels<br />

occupies ~ 60 % of the total crystal volume.<br />

Fig. 2: X-ray diffraction<br />

images of a thaumatin<br />

crystal, high-pressure<br />

cryocooled at 200 MPa. The<br />

crystal was warmed from 80<br />

to 170 K. The primary water<br />

diffuse diffraction (WDD)<br />

peak (second innermost<br />

ring) at Q = 2.03 Å -1 indicates<br />

HDA ice at 80 K, and shifts to<br />

2.00 Å -1 at 110 K, 1.92 Å -1 at<br />

140 K, and finally to 1.73 Å -1 ,<br />

indicative of LDA ice at 170<br />

K. The inner diffuse ring (Q =<br />

1.2 Å -1 ) is from an oil coating<br />

applied to the crystal.<br />

CHESS News Magazine 2009 Page 71


The radially integrated WDD profiles from a high-pressure<br />

cryocooled thaumatin crystal are shown in Figure 3.<br />

Interestingly, the 31 superimposed WDD profiles from 80<br />

to 170 K show apparent isosbestic points at Q = 2.0 and<br />

2.5 Å -1 , which suggests a possible decomposition of the<br />

intermediate states into a simple mixture of the initial<br />

and the final states. Indeed, the intermediate states could<br />

be reconstructed as a linear combination of the highdensity<br />

amorphous (HDA) ice at 80 K and the low-density<br />

amorphous (LDA) ice at 170 K. This observation provides<br />

strong evidence that the HDA ice formed inside the protein<br />

crystal undergoes a first order phase transition to LDA ice.<br />

The responses of the crystal unit cell volume and mosaicity<br />

during the phase transition are shown in Figure 4. The<br />

primary WDD peak position is superimposed as an indicator<br />

of the phase transition. A thaumatin crystal contains ~ 60<br />

% internal water and it is estimated that the crystal unitcell<br />

should expand ~ 14 % to hold all the expanded water<br />

during the phase transition based on the density of HDA<br />

ice (1.17 g/cm 3 ) and LDA ice (0.94 g/cm 3 ). Interestingly we<br />

found that the unit-cell volume expansion during the phase<br />

transition was only ~ 5 %, which is too small to hold all of<br />

the expanded water. More surprisingly, it was observed<br />

that the crystal mosaicity began to improve by ~ 25 %<br />

upon ice expansion. This indicates rearrangement of the<br />

protein molecules into better molecular packing during<br />

the ice phase transition. These observations (small unitcell<br />

expansion and unexpected mosaicity improvement)<br />

suggest that the HDA ice, which was originally in a solid<br />

state, converted to a liquid during the phase transition<br />

to LDA ice. As a control the crystal cryocooled at ambient<br />

pressure did not show the mosaicity improvement,<br />

confirming that the molecular rearrangement is not the<br />

consequence of temperature increase itself.<br />

The liquid state of water during the ice phase transition was<br />

supported by the atomic temperature factors (B-factors)<br />

shown in Figure 5. In macromolecular crystallography,<br />

the B-factors represent the effects of conformational<br />

fluctuations. The rising B-factor plots of the high-pressure<br />

cryocooled thaumatin molecule indicate that the reduced<br />

thermal motion during high-pressure cryocooling could be<br />

released during the ice phase transition, as if the molecules<br />

were exposed to a flexible environment. As a control,<br />

the B-factor of an ambient-pressure cryocooled protein<br />

molecule was monitored from 80 to 165 K and showed no<br />

significant rise.<br />

These observations have significant implications for the<br />

understanding of the anomalous properties of supercooled<br />

water, liquid water cooled below its freezing temperature.<br />

For example, the isothermal compressibility, isobaric heat<br />

capacity, and thermal expansion coefficient of supercooled<br />

water all display counterintuitive trends with decreasing<br />

temperature 6 . One promising theory that accounts for the<br />

anomalous behaviors of supercooled water is liquid-liquid<br />

(LL) critical point theory 7 . The LL critical point theory locates<br />

a second critical point of water in the supercooled region<br />

and predicts a first order phase transition between highdensity<br />

liquid (HDL) and low-density liquid (LDL) water.<br />

Fig. 3: Radially integrated WDD profiles from thaumatin crystal diffraction<br />

images. The 31 experimental WDD profiles from 80 to 170 K show apparent<br />

isosbestic points at Q = 2.0 and 2.5 Å -1 (left). WDD profiles reconstructed from<br />

2 states via singular value decomposition (SVD) analysis show a significant<br />

similarity to the experimental profiles (right). Residuals were calculated by<br />

subtracting the reconstructed profile from the experimental WDD profile at<br />

each temperature.<br />

Fig. 4: Crystal parameters from thaumatin crystals warmed from 80 to 170 K.<br />

Relative changes of the primary WDD peak position (blue circle) in d-spacing<br />

(d=2π/Q), crystal unit-cell volume (green square) and crystal mosaicity (red<br />

diamond) are shown. The values for WDD peak position and unit-cell volume<br />

are multiplied by 10. Note that the crystal high-pressure cryocooled at<br />

200 MPa shows less-than-expected unit-cell volume expansion (~5 %) and<br />

unexpected mosaicity improvement around 140 K. The crystal cryocooled<br />

at ambient pressure does not show the mosaicity improvement in the same<br />

temperature range.<br />

Fig. 5: B-factor profiles of thaumatin along its main chain. Each profile is<br />

one of 13 profiles from 80 to 165 K (blue = low temperatures and red = high<br />

temperatures). The B-factor profile from thaumatin at ambient conditions<br />

(0.1 MPa and 293 K) is superimposed as a reference (dotted black line). Note<br />

that the B-factors of the high-pressure cryocooled crystal rise dramatically<br />

during the HDA-LDA ice transition whereas the B-factors of the ambientpressure<br />

cryocooled crystal do not show significant rise in the same<br />

temperature range.<br />

Page 72 CHESS News Magazine 2009


However, experimental verification of the LL critical point theory is challenging due to spontaneous nucleation of water in<br />

the thermodynamic region in which supercooled water exists. Therefore, as an analogue to the HDL-LDL transition, the phase<br />

transition between HDA and LDA ice has been extensively studied under the assumption that HDA ice is the glassy form of HDL<br />

and LDA ice is the glassy form of LDL. However, no clear experimental evidence of the glass transition of HDA ice has thus far been<br />

presented. Furthermore, the phase transition between HDA and LDA ice has not been clearly demonstrated to be of first-order, as<br />

the intermediate states observed in this process could not be decomposed into the mixtures of HDA and LDA ice.<br />

Our experimental data from water confined in the solvent channels of protein crystals, based upon WDD profiles and Bragg<br />

diffraction analysis, suggest that HDA ice undergoes a clear first-order phase transition to LDA ice. The evidence for liquid water<br />

observed in between suggests a possible glass transition of HDA ice to HDL. Drawing upon the LL critical point theory, it is<br />

proposed that the HDA ice induced inside protein crystals by high-pressure cryocooling first transforms smoothly to HDL water<br />

(glass transition), then undergoes a first order transition to LDL water, and finally continuously converts to LDA ice through a glass<br />

transition. Further work is needed to provide more direct evidence on the existence of liquid water and to resolve subtleties in<br />

the liquid states (i.e., HDL and LDL). It also remains to be seen whether a liquid state of water exists during the HDA to LDA ice<br />

transition in the high-pressure cryocooled bulk water. These experiments may provide more insight into the phase behavior of<br />

supercooled water.<br />

References:<br />

1. C.U. Kim, B. Barstow, M.W. Tate, and S.M. Gruner; “Evidence for Liquid Water during the <strong>High</strong>-density to Low-density Amorphous Ice<br />

Transition”, Proc. Natl. Acad. Sci. USA 106, 4596-4600 (2009)<br />

2. C.U. Kim, R. Kapfer, and S.M. Gruner ; “<strong>High</strong> Pressure Cooling of Protein Crystals without Cryoprotectants”, Acta Cryst. D61,<br />

881-890 (2005)<br />

3. C.U. Kim, Q. Hao, and S.M. Gruner; “Solution of Protein Crystallographic Structures by <strong>High</strong> Pressure Cryocooling and Noble Gas<br />

Phasing”, Acta Cryst. D62, 687-694 (2006)<br />

4. C.U. Kim, Q. Hao, and S.M. Gruner; “<strong>High</strong> Pressure Cryocooling for Capillary Sample Cryoprotection and Diffraction Phasing at Long<br />

Wavelengths”, Acta Cryst. D63, 653-659 (2007)<br />

5. C.U. Kim, Y.-F. Chen, M.W. Tate, and S.M. Gruner; “Pressure Induced <strong>High</strong>-density Amorphous Ice in Protein Crystals”, J. Appl. Cryst.<br />

41, 1-7 (2008)<br />

6. P.G. Debenedetti, and H.E. Stanley; “Supercooled and Glassy Water”, Physics Today 56, 40-46 (2003)<br />

7. O. Mishima, and H.E. Stanley; “The Relationship Between Liquid, Supercooled and Glassy Water”, Nature 396, 329-335 (1998)<br />

CHESS News Magazine 2009


<strong>Cornell</strong> <strong>High</strong> <strong>Energy</strong> <strong>Synchrotron</strong> <strong>Source</strong><br />

200L Wilson Laboratory<br />

Rte 366 & Pine Tree Road<br />

Ithaca, NY 14853-8001<br />

Web: www.chess.cornell.edu

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!