You are on page 1of 153

Frontiers in Polymer Science and Engineering

Report of a 2016 NSF Workshop


{Co-sponsored by AFRL/AFOSR, ARO, DOE/BES, FDA, NIST, and ONR)
Cover images

Societal Needs
Polymers impact nearly all disciplines in science and engineering.
Inspiring the next generation of leaders will shape strategies to
address a host of grand challenges in society, including environmental
effects; energy applications; infrastructure, communications,
information; food, water, air; and health.

Macromolecular Synthesis
The development of new routes to polymeric materials is central to the
field of polymer science and engineering. Over the last half century,
important progress has been made in polymer synthesis, but rich
opportunities remain to advance precision, control, reliability, and
practical methods of synthesis.

Hierarchical Structures at Multiple Length and


Energy Scales
Polymers offer a unique opportunity to probe molecular hierarchy, from
monomeric chemistries, structures, and self-assembly at the
nanoscopic scale; to directed order over macroscopic lengths. The
ability to predict and control complex architectures remains a
formidable but increasingly achievable goal.
Integrated Measurement, Analysis, and Prediction
High-fidelity methods to characterize the structure and properties of
polymers are coupled with the advancement of the polymers field. New
experimental and computational frontiers in spectroscopy, microscopy,
scattering, and thermophysical studies will drive new materials
development.

Advancing Performance
The extraordinary and wide-ranging properties of polymers have
permeated through modern life and are indispensable to emerging
technologies. Under a unifying theme of exploring fundamental
science, research will facilitate the future discovery and development
of innovative, life-changing products.

Partnerships: Academia, Industry, and Government


Businesses, universities, and government sectors have overlapping
roles and complementary interests in advancing entrepreneurial
discovery and commercialization of polymers. Successful
collaborations require a framework for networking, vision, and
mechanisms to overcome key barriers.

Education, Diversity, Outreach


Investments in education, effective communication, inclusiveness, and
public engagement can cultivate strong future generations of leaders
in polymer science and engineering. In an evolving sociological and
technological landscape, these core values are timeless pillars of all
research communities to showcase the broader effect of polymers in
improving quality of life.
FRONTIERS IN POLYMER SCIENCE AND ENGINEERING
A National Science Foundation Sponsored Workshop
Co-sponsored by AFRL/AFOSR, ARO, DOE/BES, FDA, NIST, ONR.

Workshop Chair
• Frank S. Bates (University of Minnesota, Chemical Engineering and
Materials Science)

Co-organizers
• Patrick Brant (ExxonMobil)
• Geoffrey W. Coates (Cornell University, Chemistry)
• Jane Lipson (Dartmouth College, Chemistry)
• Chinedum Osuji (Yale University, Chemical and Environmental
Engineering)
• Juan de Pablo (University of Chicago, Institute for Molecular Engineering)
• Stuart Rowan (Case Western Reserve University; University of Chicago,
Institute for Molecular Engineering)
• Rachel Segalman (University of California at Santa Barbara, Chemical
Engineering)
• Karen I. Winey (University of Pennsylvania, Materials Science and
Engineering)

NSF Headquarters
Arlington, Virginia
August 17-19, 2016
Frontiers in Polymer Science and Engineering © 2017 by copyright holder
Published at the University of Minnesota—Twin Cities
Printed by University of Minnesota Printing Services, 2818 Como Avenue S.E., Minneapolis, MN 55414

All rights reserved. Printed in the United States of America. No part of this book may be reproduced in any manner
whatsoever without written permission except in the case of brief quotations embodied in critical articles and
reviews. For information, contact Frank Bates, Department of Chemical Engineering and Materials Science, 421
Washington Ave SE, Suite 151, Minneapolis, MN 55455.
TABLE OF CONTENTS
Participants ..................................................................................................................................... i
Invitees ......................................................................................................................................... i
Graduate Students and Post-docs ................................................................................................ ii
Special Acknowledgments .......................................................................................................... ii
Federal Agency Participants ...................................................................................................... iii

Workshop Program ..................................................................................................................... iv


Day 1—August 17, 2016 ........................................................................................................... iv
Plenary Talks .................................................................................................................. iv
Day 2—August 18, 2016 ............................................................................................................ v
Day 3—August 19, 2016 ............................................................................................................ v

Preface........................................................................................................................................... vi

Executive Summary ...................................................................................................................... 1

Section 1: Societal Needs ............................................................................................................ 12


1.1 Introduction .................................................................................................................... 12
1.1.1 Overarching Opportunities and Challenges ..................................................... 13
1.2 Environmental Effects ................................................................................................... 13
1.2.1 Major Challenges and Opportunities:............................................................... 15
1.2.2 Case Study: Food Packaging ............................................................................ 15
1.3 Energy Applications....................................................................................................... 16
1.3.1 Case Study: Batteries for Cars.......................................................................... 17
1.3.2 Major Challenges and Opportunities ................................................................ 18
1.4 Infrastructure, Communications, and Information......................................................... 19
1.4.1 Major Challenges/Opportunities: ..................................................................... 20
1.5 Water/Food/Air Applications......................................................................................... 21
1.5.1 Major Challenges and Opportunities ................................................................ 21
1.5.1.1 Water .......................................................................................................... 22
1.5.1.2 Air ............................................................................................................... 23
1.5.1.3 Food ............................................................................................................ 24
1.6 Health ............................................................................................................................. 25
1.6.1 Major Challenges and Opportunities ................................................................ 25
1.7 Aspirational, Inspirational, and Curiosity-Driven Needs .............................................. 27
1.8 Inclusive Societal Outreach and Education ................................................................... 29

Section 2: Macromolecular Synthesis ....................................................................................... 31


2.1 Introduction .................................................................................................................... 31
2.2 Significance, current state-of-the-art, and challenges .................................................... 31
2.2.1 To synthesize macromolecules of any primary structure with reasonable time
frame and cost. ................................................................................................. 31
2.2.2 To polymerize renewable monomers under the principles of ‘green chemistry’;
to create cost-competitive high-performance materials that are environmentally
benign; or to create materials that simply last forever (durable). ..................... 34
2.2.3 To develop a theoretical framework for screening polymer targets that is (i)
reliable, (ii) atomistic-scale in resolution, and (iii) efficient enough to provide
answers rapidly. ................................................................................................ 36
2.2.4 To polymerize pre-organized monomers to create programmed architectures. 38
2.2.5 To program higher order structure formation (tertiary, quaternary) into
polymerization chemistry design for in situ access to complex nano-objects. 39
2.2.6 Accelerate the development of new polymers by discovering new mechanisms
that allow preparation of new polymers from readily available monomers. .... 41
2.2.7 To develop versatile methods for sequence control and determine how
sequence affects function. ................................................................................ 43
2.3 Recommendations .......................................................................................................... 45

Section 3: Hierarchical Structures at Multiple Length Scales and Energy Scales ............... 46
3.1 Introduction .................................................................................................................... 46
3.2 Fundamental challenges ................................................................................................. 48
3.2.1 Next generation molecular and polymer building blocks for hierarchical
assembly ........................................................................................................... 49
3.2.2 Thinking of assembly as a reaction: Non-equilibrium assembly and pathways
.......................................................................................................................... 52
3.2.3 Active/Adaptive hierarchical structures ........................................................... 54
3.2.4 Bio-inspired and biological assembly .............................................................. 55
3.3 Key enabling technology needs ..................................................................................... 59
3.3.1 New tools needed for hierarchical structure and properties (synthetic,
characterization, modelling, and computation) ................................................ 59
3.3.2 Testing of properties over relevant length- and time-scales and orientations .. 60
3.3.3 For hierarchical assembly experiments: can we analyze structure without stains
or tags and within in-situ environments? ......................................................... 61
3.3.4 Core and regional facilities............................................................................... 62
3.3.5 Robust theory and modeling are integral to characterization—required to do
more detailed data analysis (dynamics from NMR, hierarchical structure
analysis, understanding of transport)................................................................ 63
3.4 Recommendations .......................................................................................................... 64

Section 4: Integrated Measurement, Analysis, and Prediction............................................... 65


4.1 Introduction .................................................................................................................... 65
4.2 Challenges, Needs, Opportunities .................................................................................. 66
4.2.1 High resolution spatial and temporal mapping of structure, dynamics, and
properties .......................................................................................................... 66
4.2.1.1 Precise characterization of monomer level structure ................................. 66
4.2.2.2 Dynamics .................................................................................................... 67
4.2.2.3 Structure characterization and chemical mapping...................................... 69
4.2.2.4 Super-resolution optical microscopy and spectroscopy ............................. 71
4.2.2.5 Unravelling complexity in structure evolution ........................................... 72
4.2.3 Advancing Structure, Property, and Event Prediction ..................................... 73
4.2.3.1 Force field development and validation ..................................................... 73
4.2.3.2 Developing and using analytical theory ..................................................... 73
4.2.3.3 Thermodynamic interactions ...................................................................... 74
4.2.3.4 Managing property data.............................................................................. 75
4.2.3.5 Predicting maximal polymer properties ..................................................... 76
4.3.1 Rare events in polymers ................................................................................... 77
4.3.2 National user facilities ...................................................................................... 78
4.3.3 Education .......................................................................................................... 79
4.4 Recommendations .......................................................................................................... 79

Section 5: Advancing Performance ........................................................................................... 81


5.1 Introduction .................................................................................................................... 81
5.2 Harnessing Process Methods to Deliver Improved Properties ...................................... 81
5.2.1 Kinetics in Complex Polymer Systems ............................................................ 82
5.2.2 Advanced Processing Methods ........................................................................ 83
5.2.3 Data-Based Methods that Include Processing Conditions ............................... 84
5.3 Control of Polymer Crystallization for Improved Performance .................................... 84
5.3.1 Bulk Polyolefins ............................................................................................... 85
5.3.2 Other Bulk Polymers ........................................................................................ 86
5.3.3 Thin Films and Confined Systems ................................................................... 87
5.4 Manipulating Interfaces and Surfaces for Combinations of Properties ......................... 88
5.4.1 Polymer-Polymer Interfaces in Polymer Mixtures and Block Copolymers ..... 88
5.4.2 Polymer-Substrate and Polymer Nanocomposite Interfaces ............................ 89
5.5 Controlling Transport in Polymers to Expand Performance Portfolio .......................... 90
5.5.1 Transport of Charge Carriers and Low Molar Mass Molecules ....................... 91
5.5.2 Transport of Waves .......................................................................................... 92
5.6 Environmental Stability, Instability, and Response ....................................................... 93
5.7 Controlling Nonlinear and Non-Equilibrium Properties................................................ 94
5.8 Recommendations .......................................................................................................... 96

Section 6: Partnerships: Academia, Industry, and Government ........................................... 98


6.1 Introduction .................................................................................................................... 98
6.2 Over-Arching Opportunities. ....................................................................................... 100
6.2.1 Changing Landscape: Evolving Challenges and Opportunities at the Academia-
Industry-Government Interface ...................................................................... 101
6.2.2 Changing Attitudes Towards Polymeric Materials and Their Environmental
Impact ............................................................................................................. 103
6.2.3 A Renewed Case for Industrial-Academic Partnerships ................................ 103
6.3 Recommendations ........................................................................................................ 104
6.3.1 IP: Intellectual Property Ownership in Sponsored Research and Technology
Transfer .......................................................................................................... 104
6.3.1.1 Collaborative Industry-Academic Research ............................................. 105
6.3.1.2 Cross-Functional Sabbaticals ................................................................... 105
6.3.2 Central Data Repository ................................................................................. 106
6.3.3 Industry-Academia-National Lab Summits.................................................... 107
6.3.4 Facilitate Greater Use of Theory and Computation in Industry ..................... 108
6.3.5 Characterization and Small-Scale Manufacturing Facilities .......................... 108
6.3.6 Postdoctoral Fellowships and Engagement Programs for Startups ................ 109
6.3.7 National Research Centers Devoted to the Advance of Polymer Science and
Engineering .................................................................................................... 109
6.3.8 Industry Supported Polymer Science and Engineering Efforts in Academia 110
6.3.9 New Scalable Routes that Reduce the Cost of Monomers ............................. 110

Section 7: Higher Education and Diversity ............................................................................ 111


7.1 Introduction .................................................................................................................. 111
7.2 Challenges in education at the university level............................................................ 111
7.3 Recommendations ........................................................................................................ 114

Section 8: Outreach and Broadening Participation............................................................... 115


8.1 Introduction .................................................................................................................. 115
8.2 Challenges and opportunities ....................................................................................... 115
8.2.1 Creating a diverse body of student scientists ................................................. 115
8.2.2 Outside the Classroom.................................................................................... 116
8.2.3 Inside the Classroom from pre-K to Grade 12 ............................................... 118
8.2.4 Network of Support and Encouragement ....................................................... 119
8.3 Public Perception of Synthetic Polymers..................................................................... 120
8.4 Recommendations ........................................................................................................ 120

References .................................................................................................................................. 122

Permissions ................................................................................................................................ 133


PARTICIPANTS
Invitees
Shenda Baker Synedgen
Anna C. Balazs University of Pittsburgh
Lisa S. Baugh ExxonMobil
Luis Campos Columbia University
Kimberly Chaffin Medtronic
Stephen Cheng University of Akron
Coray Colina University of Florida
Russell Composto University of Pennsylvania
Joseph DeSimone Carbon3D, University of North Carolina
Thomas Epps, III University of Delaware
Amalie Frischknecht Sandia National Laboratory
Venkat Ganesan University of Texas
Padma Gopalan University of Wisconsin-Madison
Peter F. Green University of Michigan; National Renewable Energy Laboratory
Robert B. Grubbs Stony Brook University
Paula T. Hammond Massachusetts Institute of Technology
Ryan Hayward University of Massachusetts, Amherst
Marc A. Hillmyer University of Minnesota
Benjamin S. Hsiao Stony Brook University
Phil Hustad Dow Chemical Company
LaShanda T.J. Korley Case Western Reserve University
Jane Lipson Dartmouth College
Timothy Lodge University of Minnesota
Christine K. Luscombe University of Washington
Jodie Lutkenhaus Texas A&M University
Krzysztof Matyjaszewski Carnegie Mellon University
Mary Ann Meador National Aeronautics and Space Administration
Bert Meijer Eindhoven University of Technology
Stephen A. Miller University of Florida
Thomas F. Miller III California Institute of Technology
Murugappan Muthukumar University of Massachusetts—Amherst
Christopher Ober Cornell University
Bradley Olsen Massachusetts Institute of Technology
Virgil Percec University of Pennsylvania
Rodney Priestley Princeton University
Richard Register Princeton University
Elsa Reichmanis Georgia Institute of Technology
Megan L. Robertson University of Houston
Thomas Russell University of Massachusetts
Anthony J. Ryan University of Sheffield
Ann B. Salamone Rochal Industries
Kenneth S. Schweizer University of Illinois—Urbana-Champaign
David Simmons University of Akron

i
Sindee Simon Texas Tech University
Christopher Soles National Institute of Standards and Technology
James C. Stevens Dow Chemical Company
Samuel Stupp Northwestern University
Brent S. Sumerlin University of Florida
Edwin L. Thomas Rice University
Matthew Tirrell University of Chicago
Carl Willis Kraton Performance Polymers
Karen Wooley Texas A&M University

Graduate Students and Post-docs


James M. Eagan Cornell University
Katie M. Greenman University of Chicago
Philip Griffin University of Pennsylvania
Joshua Lequieu University of Chicago
Yekaterina Rokhlenko Yale University
Monirosadat Sadati University of Chicago
Gabriel Sanoja University of California—Santa Barbara
Jeffrey Ting University of Minnesota and University of Chicago

Special Acknowledgments
Shrayesh Patel University of Chicago
Deborah Schneiderman University of Chicago

ii
Federal Agency Participants
Joseph Akkara National Science Foundation
Paul Armistead Dept. of Defense, Office of Naval Research
Debra J. Audus National Institute of Standards and Technology
Kathryn L. Beers National Institute of Standards and Technology
Timothy Bunning Dept. of Defense, Air Force Research Laboratory
Heather Evans National Institute of Standards and Technology
Joycelyn Harrison Dept. of Defense, Air Force Office of Scientific Research
Craig C. Henderson Office of Basic Energy Sciences, Dept. of Energy
Daryl W. Hess National Science Foundation
Irada Isayeva Office of Science and Engineering Laboratories, Food and Drug
Administration
Sean L. Jones National Science Foundation
Freddy Khoury National Science Foundation
Charles Lee Dept. of Defense, Air Force Office of Scientific Research
Eric K. Lin National Institute of Standards and Technology
Kenny Lipkowitz Dept. of Defense, Office of Naval Research
Andrew J. Lovinger National Science Foundation
Martha S. Lundberg National Institutes of Health
Mike Markowitz Office of Basic Energy Sciences, Dept. of Energy
Michael A. Meador National Nanotechnology Coordinating Office
T. Lakis Mountziaris National Science Foundation
William Olbricht National Science Foundation
Timothy Patten National Science Foundation
Dinesh Patwardhan Dept. of Health and Human Services, Food and Drug Administration
Dawanne E. Poree Dept. of Defense, Army Research Office
David Rampulla National Institutes of Health
Carole Read National Science Foundation
Linda S. Sapochak National Science Foundation
Michael Sennett Office of Basic Energy Sciences, Dept. of Energy
Alex Simonian National Science Foundation
Arthur Snow Dept. of Defense, Office of Naval Research
Suk-Wah Tam-Chang National Science Foundation
P. Thiyaga Thiyagarajan Office of Basic Energy Sciences, Dept. of Energy
Mary M. Toney National Science Foundation
Richard A. Vaia Dept. of Defense, Air Force Research Laboratory
Stephanie Watson National Institute of Standards and Technology
Timothy White Dept. of Defense, Air Force Research Laboratory

iii
WORKSHOP PROGRAM
National Science Foundation Headquarters
Arlington, VA

Day 1—August 17, 2016


7:45 AM Check-in
8:00 Refreshments (Stafford II, Room 555, 5th floor)
8:15 Introduction: Andrew Lovinger (NSF Polymers Program Director)
8:20 Welcoming Remarks: Dr. F. Fleming Crim (NSF Assistant Director for Mathematical
and Physical Sciences)
8:30 Workshop Background: Andrew Lovinger
8:40 Workshop Organization and Goals: Frank Bates (University of Minnesota)

Plenary Talks
8:45 Marc Hillmyer (U. Minnesota)—Macromolecular Synthesis
9:15 Bert Meijer (Eindhoven U. of Technology)—Hierarchical Polymeric Structures
9:45 Richard Register (Princeton U.)—Structure and Properties of Polymers
10:15 Break
10:45 Juan de Pablo (U. Chicago) / Glenn Fredrickson (UCSB/Mitsubishi)—Theory and
Simulation
11:15 Paula Hammond (MIT)—Societal Needs including energy, health and environment
11:45 Joseph DeSimone (UNC/Carbon3D)—Innovation and Translation to Practice

12:15 PM Working Lunch and Panel Discussion: Education, Diversity, Outreach


Panelists: Jane Lipson (Moderator), Shenda Baker, Coray Colina, Russ Composto,
Chinedum Osuji, Ned Thomas, Matt Tirrell

2:00 Breakout Sessions


• Macromolecular Synthesis (Topic Leader: Geoff Coates): Room 535
• Hierarchical Structures (Topic Leader: Rachel Segalman): Room 545
• Structure and Properties (Topic Leader: Chinedum Osuji): Room 555
• Advancing Performance (Topic Leader: Karen Winey): Room 565
• Societal Needs (Topic Leader: Stuart Rowan): Room 575
• Partnerships: Academia, Industry, Government (Topic Leaders: Pat Brant and
Juan de Pablo): Room 585
4:00 Break
4:30 Summary comments by Topic Leaders, feedback and discussion (Room 555)

Evening
6:00 Dinner (on your own in Topic Groups)
8:00 Continued breakout group discussions (on your own in Topic Groups)

iv
Day 2—August 18, 2016
8:00 AM Assembly and refreshments (Room 555)
8:15 General discussion (Room 555)
8:30 Final breakout session for each Topic Group (same rooms as on Day 1)
10:30 Feedback session and discussion among all (Room 555)
12:00 Lunch (on your own)
1:30 PM Breakout groups prepare summary material and begin writing of report (same rooms)
3:30 Feedback to all participants, NSF, and other Agencies
4:30 Concluding remarks and departure

Day 3—August 19, 2016


All day: Organizing committee writes Report (Main NSF building, Room 1060, 10th floor)

v
PREFACE
On August 17 and 18, 2016 a workshop was held in Arlington, VA, to assess the
opportunities and challenges facing the field of Polymer Science and Engineering. “Frontiers in
Polymer Science and Engineering”, the third decadal workshop dealing with this expansive
subject, was sponsored by the National Science Foundation (NSF) and co-sponsored by
AFRL/AFOSR, ARO, DOE/BES, FDA, NIST, and ONR. A group of 70 participants drawn from
over 50 universities, companies, and National Laboratories, focused on the most promising
directions for research associated with macromolecular materials, and on industrial, educational,
and outreach needs. This report summarizes the discussions, deliberations, and conclusions drawn
from six separate working groups that dealt with a host of technical topics. The conference
participants collectively assessed various societal challenges, which are also presented here.
As chairman, I am indebted to many colleagues and members of our professional community
for their dedicated effort in making the workshop a success and this report a reality. I am especially
grateful to Andy Lovinger (NSF), who initiated the process and provided invaluable
encouragement to me and the organizing group. Kate Beers (NIST) and Richard Vaia (AFRL)
provided very helpful insight and guidance in planning and preparing for the workshop. I am
grateful to my assistant, Phil Jensen, at the University of Minnesota for his handling of all the
logistical issues. I also single out Jeff Ting for special thanks; he began working with me on the
workshop as a graduate student at Minnesota and continued his dedicated participation in preparing
the report as a post-doctoral fellow at the University of Chicago.
Principal credit for making this project a success is directed at the remarkable group of co-
organizers who participated in shaping all aspects of this project. Pat Brant (ExxonMobil), Geoff
Coates (Cornell), Jane Lipson (Dartmouth), Chinedum Osuji (Yale), Juan de Pablo (Chicago),
Stuart Rowan (Case Western and Chicago), Rachel Segalman (Santa Barbara), and Karen Winey
(Penn) have affected every aspect of the planning and execution of the meeting and the
development and completion of this report. Their persistence in the pursuit of excellence has been
bracing. The organizers were assisted by an outstanding group of graduate students and post docs
to whom we owe our sincere appreciation. I am also indebted to the plenary speakers who most
effectively framed the discussions at the workshop.
CJ Stone has left an indelible mark on the quality of this report through his skillful and
thoughtful editing and his handling of the overall production of the document.
Finally, on behalf of the co-organizers, I offer all the participants my sincere thanks for the
inspiring insights and wisdom they injected into the workshop and the contents of this report. The
collective effort of this fantastic group of scientists and engineers has served our community well.

Frank S. Bates
Minneapolis, MN
April 18, 2017

vi
EXECUTIVE SUMMARY
Polymers are ubiquitous, affecting society in profound and extraordinary ways. The world
generates more than 300 million metric tons of synthetic polymers annually, with a market value
in excess of $400B, and this figure will grow in the foreseeable future. In the course of her life, a
newborn child will be clothed in fabrics made of polymers; transported in vehicles increasingly
reliant on advanced composites and elastomers; raised on food kept safe, sanitary, and marketable
through inexpensive plastic packaging; learn on computers manufactured using and containing a
plethora of polymeric components; drink water purified through advanced polymer-based
membrane technology; and take medicines delivered with the aid of polymer-based excipients.
Virtually all aspects of our lives are affected by advances in polymer science and engineering;
those advances continue to accelerate. This report reviews the state of knowledge and technology
of this interdisciplinary field. It assesses the challenges of and opportunities for advancing our
basic understanding of macromolecular materials. The goal is to identify promising new avenues
for advancement.
While this report deals primarily
with synthetic polymers, an important
overarching perspective is the effect of
macromolecular science beyond this
restriction. Life is based on polymers.
Cellulose and related polysaccharides,
the most pervasive polymers in the
world, provide structural support for
most forms of plant life. Animals and
insects are built from polymers of all
types, including DNA and RNA, which
store and translate genetic codes; and
proteins, which mediate innumerable
biological functions. Muscles, skin, hair,
and bones derive most of their properties
and functions from the action of
biopolymers. Most of the food we eat are
rich in polymers. Although naturally
occurring polymers are not dealt with
explicitly in this report, the scientific and engineering principles that are covered are relevant to
biologically derived macromolecules. Moreover, many of the aspirational goals identified
throughout the report are informed by the structures, properties and functions designed by nature.
The history of synthetic polymers is surprisingly short, given their effect on the global
economy. Basic concepts related to a chain architecture, advanced by Hermann Staudinger in the
1920s, gained traction throughout the 1930s and 1940s, then emerged as a dominant industry after
World War II. Breakthrough discoveries by Ziegler, Natta, and others in the 1950s catalyzed
explosive developments in polyethylene and polypropylene, today the largest segment of the
industry and still a rich source of growth and innovation. Numerous other advances in synthetic
chemistry led to the introduction of Teflon, polycarbonate, thermoset composites, high-impact
thermoplastics, adhesives, block polymers, and many other compounds that contribute to a myriad
of products. Today, polymers find unique and enabling applications across an unparalleled range

1
of markets, even within the same class of chemical structures derived from the same monomers.
For example, low density polyethylene (LDPE), one of the highest volume commodity plastics,
can be blown into film for incredibly cheap, yet highly functional, soft and flexible plastic bags.
When properly processed, ultrahigh molecular weight linear polyethylene (UHMWPE) forms
fibers with ten times the strength of steel, enabling high-value applications such as cables for
towing ships or lightweight bulletproof vests. Recent engineering advances with UHMWPE have
led to state-of-the-art artificial hip replacements that have transformed lives. And this remarkable
product diversity, all drawn from the same monomer, represents just one of thousands of types of
polymeric materials.
As with all advanced technologies, these exciting opportunities are accompanied by a host
of collateral issues. For example, society must deal with the consequences of nearly indestructible
plastics and rubbers that too often litter the landscape and pollute the oceans. Fortunately, science
and engineering driven by exploiting human ingenuity and enabled by basic and applied research
can address these challenges, as articulated here.
Polymer science and engineering is a uniquely interdisciplinary activity reliant on
fundamental principles and discoveries drawn from synthetic and physical chemistry, chemical
and materials engineering, physics, biology, applied mathematics, and other fields. Today, the
development of advanced polymeric materials to meet societal needs requires unprecedented levels
of integration across these disciplines. Work must span fundamental and applied expertise in
experimental methods, theoretical techniques, and computer simulations. Exciting grand
challenges must be envisioned and met to maintain momentum and sustain this vital component
of the world economy. Collaboration between academia, industry, and national laboratories is
essential to tackle the remarkable opportunities facing this discipline.
An NSF-sponsored workshop, “Frontiers in Polymer Science and Engineering”, was held on
August 17 and 18, 2016 in Arlington, VA, to review the state of knowledge to explore research
opportunities and challenges, to gauge related infrastructure, and to assess educational needs. The
workshop was co-sponsored by AFRL/AFOSR, ARO, DOE/BES, FDA, NIST, and ONR. It was
attended by invited workshop participants drawn from over 50 universities, companies, and
National Laboratories (see List of Organizers, Invited and Federal Agency Participants). The
program began with six plenary lectures that broadly outlined the scope of the workshop, followed
by a panel discussion with the entire group on education, diversity, and outreach. Separate working
groups then met and addressed issues related to six separate topics: (1) Societal Needs, (2)
Macromolecular Synthesis, (3) Hierarchical Structures at Multiple Length and Energy Scales, (4)
Integrated Measurement, Analysis, and Prediction, (5) Advancing Performance, and (6)
Partnerships: Academia, Industry, and Government.

Scientific and Engineering Themes


Polymers offer unparalleled opportunities for preparing materials with tailored molecular
arrangements and structures that span nano to macroscopic length scales and physical properties
ranging from liquids to high strength structural solids. Advances in synthetic chemistry have
opened a rich molecular design space ripe for exploitation. Theoretical and computer simulation
tools promise to guide and streamline the development of enhanced properties and products.
Characterizing and understanding targeted materials requires access to advanced tools, from
relatively simple and inexpensive lab devices to exceedingly sophisticated facilities maintained at
national laboratories. Exploiting this panoply of interdisciplinary scientific and engineering needs
resonated across all six working groups, resulting in the following overarching themes.

2
• Precision—Polymer synthetic chemistry has evolved to the point where previously
inconceivable macromolecular structures can be generated. We must better understand
how to bridge precise sequencing of monomers within directed macromolecular
assembly to obtain hierarchical structures with exquisitely tailored order and disorder
over length scales ranging from nanometers to microns to meters. Nature and biological
science and engineering provide particularly rich opportunities to inform the design of
synthetic polymers.
• Prediction—Expansive developments across polymer science and engineering present
exciting opportunities to create new materials with properties and applications only
dreamed of a decade ago. Yet this remarkable growth in knowledge presents a practical
dilemma, absent predictive tools to guide experimental design and data interpretation.
Significant advances in theory and computer simulation techniques are required, and they
must be supported by detailed experimental characterization of properties over the
relevant range of length- and time-scales. The results will drive predictions related to all
aspects of polymer science and engineering, from catalytic chemistry to multi-scale
modeling of molecular self-assembly and ultimate properties, such as brittle failure.
• Processing—Classical approaches to engineering new materials begin with equilibrium
concepts. Yet almost all applications of polymers rely on manufacturing methods that
exploit non-equilibrium structures, captured in useful forms through innovative
processing. Semicrystalline polyethylene, the highest-volume plastic in the world, is the
archetypical example, where chain-folded crystals interleaved with amorphous layers
generate extraordinary materials. Advances in the field are critically dependent on
modeling and manipulating polymers that are out of equilibrium.
• Properties—Synthetic polymers have gained enormous traction in society because they
offer unique access to otherwise unattainable material properties. Cutting edge research
is expanding the portfolio of commercial applications into exciting new areas, such as
flexible electronics, biomedical devices, lightweight yet impact resistant automobiles,
and energy conversion and storage technologies. Advanced composites and
nanocomposites that combine different types of plastics and elastomers, integrated with
hard or functional materials such as metals and ceramics, will open new opportunities.
• Progeny—Polymer science has inspired generations of students and professionals
through fundamental discoveries that enhance our understanding of nature and
technologies that have revolutionized modern life. The sense of awe and wonderment
that accompanies discovery and reduction of fundamental principles to engineering
practice must be conveyed to the next generation of participants through education at all
levels and be communicated to the public in clear and practical terms. Identification of
emerging problems and finding viable solutions requires diversity of critical thinking.
This demands broadening access to and inclusion of all members of our society in this
exciting field.
• Partnerships—The engine of innovation in polymer science and engineering that
emerged after World War II produced breathtaking research and development. The result
was sustained economic growth. Today, this paradigm is challenged by business models
that have largely eliminated long term research in industry. There are opportunities to
recapture the spirit of discovery with the promise of benefiting society. The creation of
jobs and the production of beneficial new technologies will be catalyzed through strategic

3
partnerships between academia, established industry, start-up companies, and
government.

Practical Effect
Advances in research carry immense practical consequences for the nation. The following
highlights emerged from the integrated considerations across all the working groups.
• Employment—Annual growth in the production of plastics and elastomers in the US is
projected to outpace the national GDP. This industry employs nearly two million people
in upstream and downstream activities that feed all aspects of the economy. Research and
development is the engine that drives innovation and growth, and it powers one of the
largest manufacturing sectors in the nation involving companies of all sizes. US
universities must meet the critical challenge of providing an educated and research-savvy
workforce to meet the nation’s employment needs.
• Environment—Synthetic polymers carry great societal benefits but also contribute to
environmental stress when improperly disposed of. Development of new plastics derived
from sustainable sources and designed to harmlessly decompose on demand or be
recycled and repurposed presents great challenges and tremendous opportunities. Basic
research into the chemistry, physics, materials science, and engineering required to meet
performance and cost objectives coupled with full life-cycle analysis can transform the
field. Cellulose, the most prevalent polymer in the world, shows how an alternative and
sustainable molecular structure could compete with polyethylene.
• Energy—Polymers affect every aspect of the energy equation, from strong and light
materials employed in windmills and automobiles to key battery components to
membranes capable of efficiently separating fuel components derived from biomass
fermentation. Buildings are insulated with polymer foams; modern energy-efficient
lighting increasingly relies on plastic electronics; and advanced fuel cells and
photovoltaic devices contain key macromolecule-based components. Fundamental
advances in polymer science and engineering are bound up with meeting the world’s
growing energy needs while reducing the environmental and societal effects of this
growth.
• Health—Enormous gains in medical science have been leveraged by exploiting the
unique properties of synthetic polymers, ranging from improving prosthetic devices to
enabling the oral delivery of insoluble pharmaceutical products to protecting against
infection. Delivering clean and safe water around the world relies on polymer-based
membrane technologies, and the growing crisis of air pollution demands new and
innovative filtration materials derived from cheap and environmentally-friendly plastics.
Every aspect of human health and the overall standard of living will benefit from progress
in the production and processing of polymers.
• Innovation—Competitive new technologies rely on insights arising from basic scientific
understanding. Research into the chemistry, structure, and properties of polymeric
materials will continue to enable development of new applications and competitive
products targeting one of the largest and most vibrant sectors of the US economy.
Collaborations among entrepreneurs, industrialists, and academicians will sustain the
idea pipeline by inspiring and spawning fundamental knowledge that underpins
innovation.

4
Overall Needs and Recommendations
Several themes and overarching concepts resonated with the workshop participants as a
whole. They are summarized here.
• Integration from Nanometers to Meters A recurrent theme was the need to connect the
development of new polymeric materials with the pursuit of synthesis, structural control,
and dynamic response across many length scales. Comprehensive structural and property
characterization across multiple length scales remains challenging, particularly in
chemically complex polymers. Multi-scale modeling from atomic to macroscopic
dimensions remains a critical limit to integrating concepts across the vast range of features
associated with polymers.
• Advanced Instrumentation Understanding and using new concepts in polymer science and
engineering is directly coupled to accessible, state-of-the art measurement tools, including
synchrotron and neutron scattering facilities at national laboratories and university-based
instrumentation, such as electron microscopy, NMR, and rheological equipment. Regional
centers should be created to provide access to academic and industry researchers and to
enhance the capabilities of national facilities. The centers should be extended to the
creation of widely-available prototype manufacturing sites for translational discoveries and
engage academia and small and large businesses.
• Ubiquitous theory and computation Once relegated to a separate category, theory and
simulation have emerged as critical tools for explaining, modeling, and predicting complex
phenomena across all aspects of polymer science and engineering. Barriers between
theoretically-focused and experimentally-rooted investigators should be further broken
down through investment in collaborative initiatives that induce exploitation of both sets
of skills. The national Materials Genome Initiative, which focuses on transforming
materials discovery so it is more collaborative and data-driven, is an example of this type
of approach. Such interdisciplinary tools could be especially significant in the corporate
setting.
• Bridging the Academic-Industry Divide The decline of basic and long-term polymer
research in industry has left a significant and dangerous void in a vital segment of the field.
Maintaining the fundamental knowledge base required for competitive innovation,
particularly in established fields, requires a new approach to collaboration between
academia and industry. This will require the development of new research funding
mechanisms, shared responsibility for managing intellectual property, and expanded
opportunities for integrating academic, industrial, and government personnel across
traditional lines of employment.

Education and Public Awareness


For the nation to remain a commanding leader in polymer science and engineering, we must
invest in education and public awareness.
• Interdisciplinary skills New synthetic polymers are designed and synthesized by creative
chemists and scaled up to industrial practice by chemical engineers. They are probed for
molecular organization, structure, and properties by materials scientists. Electrical and
biomedical engineers implement them in devices. Theoreticians steeped in applied
mathematics and physics model them at all length scales. Few other areas of science and
technology require and embrace such a breadth of knowledge. Practitioners face the
exhilaration of interdisciplinary discovery while being challenged by a daunting array of

5
concepts and required technical tools. New approaches are essential to providing students
and industrial scientists and engineers with the interdisciplinary skills required to advance
research and development of polymeric materials.
• Broadening workforce diversity Rewarding opportunities will be part of a more diverse
workforce in polymer science and engineering. We must confront traditional biases in
hiring and career development. This includes dispelling various misconceptions and
changing attitudes and expectations across all age groups. We dare not miss the chance to
enrich our lives with the unpredictable creative contributions that come with diversity of
thought, culture, and intellectual ambition.
• Public engagement Public perception of polymers is shaped by relatively few, often ill-
informed concepts devoid of a central appreciation of science and engineering.
Surprisingly few people get how macromolecular concepts are woven into daily living,
forming the basic ingredients of life and contributing to everyday health care and nutritious
food. Plastics have contributed innumerably to improved standards of living around the
world, and virtually limitless opportunities await discovery and development. We must
share with society the sense of awe and excitement we experience when faced with the
marvels of advances in the science and engineering of macromolecular materials. We must
share as a community, across all disciplines, and with engagement by industry,
government, and academia.

Working Group Recommendations


Societal Needs
• Environmental Effects: It is important to redefine the relationship between the
environment and polymer science and industry. Thus, a key goal is the development of
accessible, scalable polymers that match or exceed the property matrix of existing materials
while having a greener life-cycle. Finding more efficient ways to recycle, e.g., without the
need to separate polymers, is also a key challenge.
• Energy Applications: The ever-increasing energy requirements of a growing population
will likely be met by a combination of generation and storage. Society needs polymers that
can improve the efficiency of renewable energy processes (e.g., wind, solar, and fuel cells),
drop their cost, and improve the capacity and safety of energy storage systems.
• Infrastructure, Communications, and Information: Mass migrations and a growing
population require rapid, low-cost, local production of lightweight, robust infrastructure.
Thus, we need to develop new integrated materials and smart technology that provide
durable, healthy, environmentally-conscious, aesthetically-pleasing, lightweight, modular,
and inexpensive building materials to replace our aging infrastructure and meet the needs
of a rising global population. Smart construction materials that improve energy efficiency
and transport clean water are areas polymer technology can have a significant effect.
• Food, Water, and Air: A new generation of materials is required if we are to meet the
food, water, and air needs of a growing population. Polymers can address these key
challenges including membranes for water and air purification; pollution sensing
technologies; more efficient ways to transport water; and environmentally-friendly
packaging materials that lengthen the shelf life of food.
• Health: Polymers will continue to be critical to improving human health. Innovative
materials are needed to control, treat, and diagnose infectious diseases cost-effectively. A
better understanding of how polymers react in harsh and dynamic environments in the body

6
will significantly affect implants, multi-functional biosensing, and drug delivery systems
and open the door to new polymer-based technologies in vivo.
• Aspirational, Inspirational, and Curiosity-Driven Needs: US leadership in polymer
science and engineering feeds societal progress and global economic development. US
leadership must foster inspirational, curiosity-driven research and discovery. This is tightly
coupled to aspirations to unlock the secrets of nature, which often spawn intellectual
environments that support the development of disruptive and game-changing technologies.

Macromolecular Synthesis
• Synthetic Methods: An overarching challenge is the development of robust, easily-
implemented methods to synthesize a wide range of macromolecules from readily available
monomers. Better control over sequence, composition, tacticity, branching, architecture,
topology, dimensionality, functionality, and dispersity is needed. Further, the development
of new mechanisms of polymerization and combinations of existing mechanisms will
afford new materials from existing and prevalent monomers. Addressing these challenges
will provide novel polymers that significantly affect the ability to answer fundamental
questions in polymer science.
• Green Polymers: We need to consider the principles of “green chemistry” in polymer
synthesis and create cost-competitive, high-performance materials. Strategies that improve
recyclability, reduce environmental effects, and control the durability of polymers can
significantly change the way society uses plastics.
• Screening of Theoretical Targets: It is important to establish the theoretical framework
for screening polymer targets on rapid timescales. The methods should be (i) reliable for a
diversity of properties, (ii) atomistic-scale in resolution to account for the effects of subtle
structural changes, and (iii) efficient enough to explore fully the chemical space.
• Polymer Assemblies: We need to program higher-order structure formation into
polymerization chemistry design for in situ access to complex nano-objects and polymerize
pre-organized monomers to create programmed architectures. Synthetic methods that
directly afford higher-order structures will open the door to new technologies and
applications of macromolecules.

Hierarchical Structures at Multiple Length and Energy Scales


• Complex Assembly: We must learn to direct the assembly of inherently complex and
multi-functional building blocks. This includes retaining the biological functionality that
results from hierarchical structure in a designed system. We can currently make a
macroscopic heart valve that retains its macroscopic functionality but not the mesoscopic
functionality of a cell or the microscopic functionality of enzymes. Similarly, the transfer
of chirality and other attributes from the molecular length scale to the macroscopic
presents largely-untapped opportunities to control structure and properties.
• Using Hierarchical Energy Scales to Control Assembly: Unique opportunities include
the rational design of polymers with macroscopic properties that emerge from
hierarchical structures resulting from tradeoffs between short and long range non-
covalent interactions. Non-covalent bonding (e.g., electrostatic, hydrophobic, hydrogen)
is critical for developing stimuli-responsive, adaptive, and reversible materials. The
interaction energies dictate the hierarchical structures and respond to multiple orthogonal

7
handles on faster timescales than their covalent counterparts. In addition, the chain
connectivity results in long-ranged topological correlations in space and time. We must
learn to precisely specify the interplay between polymer connectivity and segmental
interaction energy to gain control over hierarchical structure and properties.
• Self-Assembly in 3D: Directed self-assembly techniques have made great strides in
lithographic applications, but further advancements require us to learn to direct assembly
in three dimensions. As integrated circuit design bridges to three dimensions, lithography
is a critical enabling technology. Further, patterning in three dimensions opens the
possibility of designing bulk material properties (tuned and directed conduction of ions,
electrons, or molecules; or highly directional mechanical properties) or integration with
biological materials.
• Dynamic Structures: Hierarchically engineered materials are treated as if their structure
is static. The ability to pattern dynamic architectures to make materials that can change
properties on demand or sense and heal their own defects is a great opportunity. This will
enable new paradigms in the design of materials for switches, membranes, electronics,
biological materials, and structural components.
• Structural Pathways: We must begin to think of assembly not as a process with fixed
thermodynamic states defined by equilibrium but rather as a pathway with useful
intermediates that yields a richer and more complex phase space. Assembly can be
designed in a manner similar to a chemical reaction path, where many unit operations
simultaneously take advantage of the kinetics and thermodynamics of assembly.

Integrated Measurement, Analysis, and Prediction


• Measurement Tools: Addressing the grand challenges in polymer science requires the
elucidation of structure and properties with high fidelity at ever smaller length scales in a
broadening range of environments, including manufacturing-relevant situations. We must
develop new tools and methods to characterize structure and property with the fidelity and
resolutions required to significantly advance current knowledge. Concurrently, more
effective use of existing and emerging advanced experimental tools and methods to address
open questions in polymer science should be encouraged.
• Analysis: The analysis of structure and property data presents unique challenges in terms
of handling data and extracting useful information. The need for multi-modal and real-time
measurements in complex environments is increasing, and it is intensifying these
challenges. At the same time, the most routinely-used analysis methods are out of date.
Strategies are needed to accelerate dissemination of up-to-date theory and algorithms for
data analysis; and to make more effective use of the multiple large data streams provided
by today’s experiments. NSF and other federal agency support for the development and
implementation of such strategies is recommended.
• Data: Developing new polymers and evaluating new theoretical and computational
approaches rely critically on the availability of high-quality thermophysical and other
property data. There is a pressing need for accessible and well-managed cyberinfrastructure
that facilitates and encourages the storage, organization, curation, retrieval, and effective
use of experimental and computational data on polymer properties. Federal funding
agencies should support efforts to address this need. Providing such infrastructure and
associated frameworks would help accelerate the discovery of new, useful polymeric
materials.

8
• Prediction: Accurate and reliable prediction of structure and properties(including maximal
properties of polymeric materials and the onset of failure due to rare events) would be a
transformative advance. Mechanisms should be provided to intensify theory, simulation,
and experimental efforts to further first principles prediction.
• National Facilities: Access to specialized capabilities at national laboratories, including
midscale instrumentation, is critical to advancing polymer science. Likewise, it is vitally
important for potential users to have access to the training required to fully exploit
facilities’ unique capabilities. The community should work closely with facility personnel
to ensure users’ needs are met, particularly when planning future capabilities. Further,
consistent and high-quality geographically-distributed user training in advanced
experimental and computational techniques should be ensured to maximize the effect of
science conducted using national and regional facilities.

Advancing Performance
• Out of Equilibrium: Polymer processing frequently employs far-from-equilibrium
conditions and non-linear deformations such that the design of processing methods to
control polymer properties is often accomplished by good intuition and empirical findings.
There has been considerable progress correlating processing condition to polymer
structure, most notably in semi-crystalline polymers, but the connection to advanced
properties is less-well developed. Experimentally-robust correlations and predictive
theories, simulations, and data bases, should be built so current and emerging processing
methods can be manipulated to control multi-faceted polymer performance. Particular
opportunities exist for polymers containing nanoparticles, for polymers that possess
distinct chemical moieties that produce nanoscale morphologies, and for semi-crystalline
polymers.
• Interfaces: Polymers are nanoscale molecules and are often used under confinement,
including thin films and polymer nanocomposites. The interfaces and surfaces that impose
this confinement are known to influence properties and performance, although the
compositional profiles and the polymer conformations are better understood than the effect
on macroscopic properties. To fully optimize polymer performance, the role of interfaces
and surfaces on linear and non-linear properties needs to be developed.
• Transport Properties: The importance of molecular, proton, ion, and wave transport
through polymeric materials has grown in the last decade. There are urgent needs for
improved membranes, polymer electrolytes, advanced functional polymers, sound
damping materials, and many others. Developing polymers with tunable transport
properties requires additional understanding of the underlying fundamental mechanisms
associated with the various driving forces.
• Non-linear Response: The chemical and architectural diversity available in
macromolecules provides abundant opportunities to engineer materials that respond to
stimuli. Technologies have yet to exploit the full array of linear and non-linear mechanical
properties, chemically-triggered responses, reversible responses, and resistance to response
in extreme environments. Exploiting these behaviors in advanced applications requires
developing predictive theory, engineering models, and integrated experiments to capture
the associated multiscale phenomena.

9
Partnerships: Academia, Industry, and Government
• Intellectual Property Agreements: Complex intellectual property negotiations can often
hinder the development of partnerships. It is essential to find solutions that facilitate
productive interactions. We recommend federal agencies convene a panel of funded
scientists and university and industrial tech transfer experts to assess best practices in
sponsored research. They should make recommendations and provide a series of
reasonable IP template(s) to reduce the time and energy required to implement industrial-
academic research partnerships, thereby increasing their likelihood of success.
• Enhanced Industry-Academia Interactions: Entrepreneurialism has become an
important source of innovation for the US polymer industry. Pathways should be identified
to create more effective communication between industry and academic partners, which (i)
explain recent academic breakthroughs to industry and (ii) tell academia about critical
industry needs. Mechanisms for collaborative research efforts—which could include joint
research centers, shared-instrument facilities, or joint symposia—are critical to maintain
these relationships. It is recommended that NSF or other federal agencies host Industrial-
Academic-National Laboratory summits. The summits would raise awareness and
excitement around research topics that have the scope and the scale to affect the entire
scientific community and change the direction and speed of innovation.
• Industrial Sabbaticals: A new sabbatical program should be created, co-sponsored by
NSF and other federal agencies, to let agency-supported researchers spend extended
periods in industry or at a national lab. Such a program could be expanded to include
jointly-advised postdoctoral fellows so access to world-class instrumentation is more
readily available to aid research addressing grand challenges in polymer science and
engineering. Programs that rotate graduate students through industry labs and expose them
to industrial needs would be extremely beneficial; existing paths should be improved by
creating new paths that let graduate students work in industry for an appropriate time.
• Industrial-Academic Partnerships in Computation: Industry, particularly small and
startup companies, does not generally have access to the theoretical and computational
expertise and the infrastructure available at national labs and at universities. Efforts are
increasing to advance education in computer science and big data management. Facilitating
more effective communication can bridge gaps between experimentalists and
computational personnel. We should create mechanisms to tell industrial researchers about
the theoretical/computational expertise available and to increase their access to it.
• Polymer Research Centers: Multiple major research and development centers should be
established, with support by NSF and other federal agencies, to foster continued innovation
and to secure global leadership in the US polymer industry. Such centers should address
problems across a broad spectrum of fundamental and applied research areas including
synthesis, characterization, and advanced computational modeling. These unique centers
should be embedded within universities, thereby providing the highly-skilled technical
expertise and infrastructure necessary to compete in industry on a global scale.

Higher Education and Diversity


• Training: The number of academic institutions offering undergraduate majors in polymer
science and engineering should be increased. The depth and breadth of offerings in those
that do should be increased. Masters and Doctoral programs should be deepened and
broadened, too. The first step is to identify an optimal set of core elements for each degree.

10
This will let institutions acquire resources needed to ensure students graduate with broad
backgrounds, targeted expertise, and strong critical thinking skills.
• Curricula: Not all academic institutions have the faculty resources to let them offer the
degree programs described in the first objective. Thus, another key goal is the creation of
new resources to implement the curricula, especially for institutions lacking scientific
expertise. Activities that target the core curricular need include workshops, online modules,
and summer schools.
• Diverse Workforce: Solving the most difficult problems in polymer science and
engineering will require attracting and retaining a diverse workforce of critical thinkers
with strong backgrounds. However, bias in recruiting and advancement remains an issue.
Active strategies are needed to counter this; ensuring best practices are being followed
requires localizing resources to address these issues.

Outreach and Broadening Participation


• Public Awareness: Exposure to the excitement and potential of polymer science and
engineering at an early age is important for encouraging future scientists and educating all
citizens. Rather than relying on the interests of particular educators, there should be a
website for those who want to develop curricular initiatives incorporating polymer science
and engineering. To make it easy to connect with a state's science curriculum, this site
should include links to core K-12 requirements for each state.
• Direct Access: Hands-on experiments highlighting polymer science and engineering have
the greatest potential for engaging and exciting children at all grade levels, but adapting
experiments and implementing them involves effort not feasible for most teachers. Student-
ready experiments-in-a-box (“Blue Labcoat”) compatible with core requirements should
be developed. Implementing them should not require any school-funded special training or
resources.
• Public Engagement: School should not be the only forum for generating excitement about
polymer science and engineering. A powerful reinforcement to school activities would be
to involve childrens’ families. There needs to be a broad and effective set of programs and
activities for scientists and engineers to engage the public where they gather.

11
SECTION 1: SOCIETAL NEEDS
Discussion Leader
Stuart J. Rowan, Case Western Reserve University and University of Chicago

Group Members
Peter F. Green, University of Michigan; National Renewable Energy
Laboratory
Paula T. Hammond, Massachusetts Institute of Technology
Benjamin S. Hsiao, Stony Brook University
LaShanda T.J. Korley, Case Western Reserve University
Christine K. Luscombe, University of Washington
Stephen A. Miller, University of Florida
Megan L. Robertson, University of Houston
Ann B. Salamone, President of Rochal Industries
Kenneth S. Schweizer, University of Illinois Urbana-Champaign

Special Acknowledgements
Katie M. Greenman, University of Chicago, Student Participant and Contributor
to this report
Shrayesh Patel, University of Chicago, Contributor to this report
Deborah Schneiderman, University of Chicago, Contributor to this report
Jeff Ting, University of Chicago, Student Participant and Contributor to this
report

1.1 Introduction
From the fabrics that make our clothes, to the plastics and foams that constitute our furniture
and shoes, to their ubiquitous presence in lightweight food and drink packaging, there is hardly
any aspect of our life today that is not affected and improved by polymers. Plastics and elastomers
play key roles in biomedical technology (e.g., contact lenses, dental implants, prosthetics, blood
bags) and in the electronics industry (lithographic technology, plastic electronics, computer
housing). They hold products together (glues and adhesives), play critical roles in building
construction (from vinyl sidings to insulating foam), are found in shower gels and motor oil (as
viscosity modifiers), and in food itself. Airplanes (the Boeing 787 Dreamliner) and automobiles
increasingly contain greater amounts of and more sophisticated types of plastics and composites.
Polymer science and the associated industry have played a dramatic role in how we live in the 21st
century, and it would be fair to characterize this era as the “plastic age”; thus, we should ask: what
are the societal needs and opportunities facing polymer science and engineering over the next
decade and beyond?
Identifying societal needs is one way to help harness the imagination of scientists, who must
make breakthroughs to keep pace with the problems that beset our world. For example, it is
predicted enough water, food, and energy will be needed to sustain 9-10 billion people in 2050.[1]

12
In addition to rapid population growth, other societal challenges include: strain on natural
resources, climate change, environmental pollution, proliferation of antibiotic-resistant diseases,
infectious diseases, and mass migration. It is likely a combination of legislative solutions and
technical solutions will be needed. In many cases, technical solutions will rely on the development
of advanced materials to access property profiles that do not exist. The future of the polymer field
lies in identifying and designing these new, possibly novel, materials. Success will require
cooperation between academic and industry researchers, non-government entities, and legislatures.
This chapter focuses on identifying and describing seven key overarching opportunities and
challenges driven by societal needs. The goal is to outline areas where advanced polymer research
and technology will have an impact. Possible solutions to these challenges are introduced briefly
here and discussed in greater detail in the following sections of this report.

1.1.1 Overarching Opportunities and Challenges


Seven areas of overarching opportunities and challenges that have been identified are:

• Environmental Impacts
• Energy Applications
• Infrastructure, Communications, and Information
• Food/Water/Air
• Health
• Aspirational-, Inspirational-, and Curiosity-Driven Needs
• Inclusive Societal Outreach and Education

Each of these is considered below, including identification of Grand Challenges and a


number of outstanding opportunities.

1.2 Environmental Effects


One critical area of research in the next decade and beyond will involve changing the
relationship between polymers and the environment. This objective encompasses a wide range of
science and engineering disciplines, ranging from basic polymer science (synthesis, property
characterization/understanding and polymer degradation) to environmental studies (including the
bioevaluation of the monomers, polymers, and polymer-degradation products).
Grand Challenge: Achieve accessible, scalable polymers that match or exceed the
property matrix of existing materials, yet have a green life-cycle.
A critical aspect of this field is the evolving area of green polymers. Currently, the term
“green polymer/plastic” is used to cover a multitude of different issues, ranging from a material
made from renewable resources or requiring less energy for production to those that are
biodegradable or recyclable. A material made from a renewable resource is often not recyclable or
biodegradable. It is important to differentiate the specific traits of a “green polymer”, by addressing
the life span of a given polymer (Figure 1.1) from birth (synthesis); through life (being used in its
desired application); to death or rebirth (what happens once it is no longer used for its desired
application). Each of these comes with challenges, some linked. Of course, the ideal polymer is
green in all three areas but that may not always be possible, necessary, or desired.
Green feedstocks, such as (bio)renewable and under-utilized waste for monomers and
polymers (i.e., a green birth), are important. The use of biopolymers, such as cellulose, natural

13
rubber and natural polypeptides (e.g., wool
and silk), have played a significant role in
the development of the field.[2] The
chemical modification of such polymers
(e.g., nitration or methylation of cellulose)
allowed access to new property profiles
and expanded the utility of such materials.
The emergence of oil and gas as cheap and
readily available raw materials resulted in
dramatic improvements in the cost
Figure 1.1. Reconstructing the lifecycle of plastics to effectiveness and manufacture of a wide
include a green birth, green life, and green death/rebirth variety of polymers. They could be tailored
improves the environmental impact of polymers.
to have myriad desirable properties, e.g.,
rigid, soft, elastic, transparent,
impermeable, insulating, bioactive, etc. Oil/gas-based polymers are currently the majority of
commercial polymeric materials,[3] but synthetic polymers from biorenewable monomers (e.g.,
poly(lactide) and polyhydroxyalkanoates) are gaining traction. A key issue is to improve the range
of physical properties accessed by such renewable polymers. For instance, aromatic compounds
from biosources, such as lignin (a major structural component of wood), can be employed directly
or converted into monomers using chemical transformations. A different approach is identifying
and investigating new types of materials, such as nanocellulose and nanochitin from biosources.[4]
One issue associated with some biorenewable polymers is biosources that compete with food
production. A potential strategy is to use waste or bio-waste as a feedstock. A 2013 estimate found
the US alone produced 550 million dry tons of crop waste residue annually, of which 70% is not
reused. An alternative is to look to the estimated 467 million dry tons of non-food crops (e.g.,
Miscanthus x. Giganteus or switch grass) [2] that can grow in abandoned croplands or areas not
suitable for food crops and, as such, do not complete with food production.
Evaluation and reduction of the ecological effect of polymers after use to ensure a green
death or rebirth are different but equally important aspects of green polymers. Strategies include:
development of biodegradable polymers and enhanced recycling processes, such as handling
mixtures/heterogeneities; incorporating degradation processes into recycling; promoting more
efficient sorting of plastics; and extending the range of recyclable polymers to non-traditional
plastics, thermosets, elastomers, and other material classes.
A full life cycle analysis (LCA) must be developed and implemented to improve evaluation
of real-life sustainability metrics. Such metrics are necessary for a complete understanding of the
environmental effect of the production and the use of a specific polymer. These LCAs take into
account all aspects of polymer development, production, implementation, and destruction, creating
a comprehensive assessment of the effect a polymer has on the environment, regardless of its
under- or over-performance at a particular stage of its life. For example, polyolefins produced from
oil and gas are generally not considered green polymers, but they score well in LCAs, in part as a
consequence of their high atom economy and low pollutant emissions during synthesis and
processing.[5] Polyethylene and polypropylene offer innumerable benefits to society, and the
scientific and engineering community is obligated to find new ways to recycle and reuse these
extraordinary materials.
Production of polymers and plastics on a large scale raises important environmental and
health concerns. Recent attention on the use of bisphenol A and phthalates as monomers/additives

14
highlights the need for rigorous and comprehensive assessment of toxicity. Polymers, additives,
monomers, and degradation products must not harm consumers, even unintentionally.
Concomitant with this the community must find non-toxic/benign environmental replacements for
potentially harmful systems that still allow access to the desired property profiles.
Cost and availability of raw materials cannot be ignored. Generic to all phases of the
materials life cycle are the resources (energy, water, cost, etc.) required to synthesize, process, or
recycle polymers. If producing commodity polymers achieves a dramatic reduction in energy and
resources, favorable effects can be realized. Integrating big data with LCA is an important step to
achieving a holistic approach.

1.2.1 Major Challenges and Opportunities:


Green Birth. The focus in this area is on the raw materials and synthetic methods used to
make polymers. Obtaining monomers/polymers from renewable biomass-based resources has been
emphasized in the past decade. This focus is still a critical area of research, but additional attention
is needed to access competitive polymer properties via scalable and cost-effective materials and
processes. This includes exploiting known bioderived monomers or finding new ones; improving
methods in biorefineries (e.g., to access new monomers or improve yields on a large scale); and
investigating scalable microbe polymer synthesis. The global surplus of raw materials offers an
alternative route to polymer Green Birth. This surplus includes: agricultural plant waste; industrial
waste byproducts (e.g., sulfur production as a byproduct of the oil industry); and waste from human
consumption (i.e., efficiently harvesting ocean plastics or landfill sites as potential raw materials).
Underused yet naturally-abundant feedstocks not currently thought of as good raw materials could
be considered, e.g., sand, limestone, and carbon dioxide.
Green Life. The focus here is on how the material behaves during its functional lifetime. It
includes extending the lifetime and durability, although designing a material that is durable and
degradable is a fundamental challenge. Healable materials with enhanced property profiles may
provide some answers. Understanding polymer aging during use is important, with the goal of
developing (and extending) the full-property lifetime. Toxic or environmentally harmful
byproducts must be better understood, for example, through monomer/additive leaching or
degradation products. Replacing known toxic compounds (such as BPA) with benign compounds
that meet (or ideally exceed) current performance is a key.
Green Death/Rebirth. Green death is focused on how the polymer (bio)degrades; green
rebirth is focused on recycling or reuse. From the green death perspective, an ideal material
maintains superior performance for its required use, but degrades efficiently and quickly on
demand. Therefore, much attention should be paid to developing materials that can degrade under
targeted conditions (e.g., the degradation profile of a polymer in a desert is quite different from
that of an ocean). The degradation rate should be at least equal to—ideally, greater than—the rate
at which the waste products are generated. Of course, there are numerous applications where the
material is required to last for a very long time (e.g., polymers used in infrastructure), so not every
material needs to degrade quickly; but there is still a need to think about what happens to durable,
longer-lasting materials after their use is over. About 28% of all plastic is collected for further use;
about half of that is incinerated for energy, and the rest is recycled.

1.2.2 Case Study: Food Packaging


Plastic packaging offers a very visible example of what can happen if a green death/rebirth
is not fully considered. Food packaging accounts for 26% of the total volume of plastics used today
(see Section 1.5., Water Food Air Applications, for more information on food packaging).

15
Polymers are essential to the safe and reliable marketing of agricultural products. But the enormous
benefits are accompanied by significant concerns. According to the 2016 report by the Ellen
MacArthur Foundation, entitled “The New Plastics Economy: Rethinking the Future of Plastics”,
the annual worldwide production of plastic packaging is over 80 million tonnes. [3] 40% of this
ends up in landfills; 32% is “leaked” into the environment, both land and sea. At least 8 million
tonnes of plastic packaging end up in the oceans every year. If this continues, by 2050 there will
be more plastic by weight in the oceans than fish! It is therefore imperative that polymers used for
packaging be made with significant consideration focused on environmental effect, both in
production and after use. The recycling of plastic packaging is much lower (14%) than that for
other common packaging materials, such as paper (58%) or iron/steel (70-90%). Furthermore,
while plastic packaging can be a highly-engineered product (for example, high barrier plastics are
complex multi-layered materials), recycled plastic is generally used for products with much lower
value. Thus, not only is there an environmental effect to the limited recycling of plastic, there is
also a significant economic effect, too. It is estimated the original packaging retains only about 5%
of its worth after accounting for limited recycling, the lower-value products, and value losses in
sorting and reprocessing. As such, estimates suggest the loss to the economy can be as high as
$120 million/year.[3] To improve green rebirth, more efficient and widespread recycling processes
must be developed. Most recycled thermoplastics result in materials with lower property profiles.
A key goal is upcycling: recycled materials that exhibit even better property profiles than the
original polymers. Furthermore, recycling processes should be extended to polymers that are
traditionally not degradable or recyclable, such as thermosets and elastomers. One of the major
challenges in recycling polymers is separating the different types of polymers. Thus, concurrent
with upcycling and expanding recycling, reinventing facile separation processes or creating
recycling without separation would have a significant effect on the green rebirth of polymers.

1.3 Energy Applications


An increasing human population with increasing standards of living, especially in developing
regions, is accelerating energy consumption. A 2016 report on the international energy outlook
predicts energy demands across the world will increase by almost 50% by 2040, with most of this
increase coming from developing countries.[6] The augmented infrastructure—electrification,
transportation—will require significant enhancements in energy generation, storage, and
distribution. The environmental consequences—extreme temperature/weather events, rising sea
levels, fresh water availability, disruption of ecosystems—pose additional challenges if these
energy needs are met primarily by fossil fuels. The central and urgent issue is to reduce human-
generated greenhouse gas (GHG) and other harmful emissions. For example, the atmospheric
concentration of CO 2 is already two-thirds of the GHGs required to get to the 2 °C “limit” set by
the Paris agreement. Deep decarbonization pathway analyses propose that any successful plan
must include three basic elements: energy efficiency, decarbonized electricity, and end-use fuel
switching to electric sources.[7] Specific measures mentioned include: targets for renewable
energy; nuclear energy; a more sophisticated power distribution grid that can manage a diverse
range of storage and generation sources; low-emission agriculture and waste management; and a
reduction in deforestation. Polymers will play a significant role in all aspects of emerging solutions
to these challenges.
Fundamental scientific discoveries leading to improved efficiency and cost are required for:
renewable energy; innovative and energy-efficient building designs; energy storage; alternative
low-carbon transport fuels (e.g., biofuels and hydrogen); agriculture and forestry bio-based

16
chemicals and materials; and theoretical and computational models that inform science. It is likely
the power grid of the future will need to manage a diverse range of power sources (geothermal,
biomass, solar, wind, hydro, nuclear, natural gas, coal, etc.) and storage devices. There are obvious
and undiscovered opportunities for polymers and polymer composites to play a significant role.
New battery technologies, beyond the current lithium-ion technologies, are a prime example of an
urgent need. New batteries should be safe, cheap, efficient, and have high storage capacities with
long reliability. A deeper understanding of storage and transport of waste heat from buildings and
industrial processes is also required. In solar energy, new polymer-based materials and
manufacturing processes will lead to lower cost/scalable solar modules and panels. In wind energy,
new materials are needed that meet criteria such as lightweight, low-cost, reliable, processable,
durable, and resistant to aging. These challenges need to be met if the goals of the wind vision
study are to be achieved: 10% of US energy demand generated by wind by 2020 (113 gigawatts,
GW), 20% by 2030 (224 GW), and 35% by 2050 (404 GW).[8] In all energy generation and storage
technologies, life cycle analyses and manufacturability need to be considered. And all these
technology goals include extensive use of advanced engineered polymers.
In general, a low-carbon energy economy provides new opportunities for scientific discovery
and innovation essential for societal and economic progress around the globe.
Grand Challenge: Energy storage and generation materials with a combined 10x
improvement in cost, properties, and performance
The development of new materials will be critical in the transition to a more diverse energy
portfolio. This includes new polymers for power generation, (e.g., fuel cell membranes,
composites for wind turbines, batteries and super capacitors for energy storage) and lightweight
materials to reduce transportation energy consumption and costs. The advantages of polymers in
energy-based applications include low cost and comparative ease of processing. However, a new
generation of polymers and polymer-based composites with new performance characteristics will
be necessary to meet these energy challenges Developing polymers that resist aging and are
durable for long periods under diverse conditions are key goals, e.g., for encapsulation of
photovoltaic devices. Life-time analysis of energy-related materials (degradation, stability,
performance, etc.) is critical for translation into the real world. New predictive theoretical and
computational tools will be essential.

1.3.1 Case Study: Batteries for Cars


We consider energy storage requirements for a vehicle. The current lithium-ion battery
technology has a practical specific energy of 120-150 Wh/kg and energy density of 250 Wh/L.[9]
Depending on the total weight and volume of the battery module and the vehicle design, the driving
range in electric vehicles is around 100-300 miles. To increase the driving range while also
reducing the battery module size, weight, and cost, it is necessary to increase the specific energy
and the energy density to above 350 Wh/kg and 750 Wh/L, respectively.[10] While advancing
these technologies, one must also pay careful attention to the inherent safety of the battery. From
the Boeing Dreamliner to the Galaxy Note 7, issues of lithium-ion battery safety have hit the
headlines. Herein lies the need for functional polymeric materials where ionic, electronic, and
redox-active properties can be leveraged as inherently safer transporting materials and even as
electrodes in new battery technologies.
A critical application of polymers is as a membrane separator between anode and cathode.
These membranes must be mechanically robust to prevent damage during operation; ion
conducting and stable in both oxidative and reductive environments; and stable over multiple

17
charge/discharge cycles. Most current polymer separators have been developed for use with
lithium-ion batteries; however, there is a need to develop membranes for other classes of batteries.
For example, magnesium ions are doubly charged, which translates to nearly double the volumetric
energy density compared to lithium metal anodes (~4800 and ~2000 Ah/L, respectively).[11] This
need challenges the polymer scientist to create polymer electrolytes that are both good magnesium
ion conductors and have a large electrochemical stability window. Alternatively, sulfur is another
promising cathode material, with a specific capacity of 1600 Ah/kg and an operating voltage near
2 V vs. Li/Li+.[9] When coupled with a lithium-metal anode, sulfur batteries have a theoretical
specific energy of 2500 Wh/kg and a practical range of 350 Wh/kg to 600 Wh/kg.[9] A challenge
for the sulfur battery is the dissolution and shuttling of intermediate polysulfides during the
charge/discharge process, which significantly reduces cycle life. Polymers could be used to either
prevent the dissolution of polysulfide salts or act as membranes that block the transport of
polysulfide salts.

1.3.2 Major Challenges and Opportunities


The energy sector is replete with challenges and opportunities that can be uniquely addressed
by advanced polymer science and engineering. These include:
• Biomass-based electronic materials: Naturally-occurring conjugated macromolecules
(such as chlorophyll, indigo, and natural dyes like melanin) have largely untapped
potential as active components in electrical applications ranging from energy harvesting
to biomedical-electrical interfaces.[12]
• New materials for thermo-electric applications: Organic thermoelectric devices based on
doped semiconducting polymers are of great interest because of the potential for
developing low-cost thermal harvesters of low-grade heat (<200 °C). They could power
wearable devices and allow local temperature control.
• All-organic batteries (without metal electrodes) either redox-flow batteries or as solid
state polymer-based batteries.[13, 14] The development and understanding of redox
active polymers is a key components of this goal.
• New materials for lightweight batteries, solar panels, wind-blades, transportation: The
materials directly involved in the generation, delivery, or storage of energy are important,
but the infrastructure surrounding the energetic material must be equally sophisticated.
• Smart materials for multi-tasking, multi-functional energy harvesting/storage: With the
implementation of new sources of energy (wind, solar, geothermal, etc.), towns, cities,
and countries will require updated resources for distributing and storing the energy.
Developing smart polymeric materials that can be seamlessly incorporated into grid
systems as sensors would significantly enhance the ability to transport energy to
consumers.
• Personal energy generation: The lightweight and pliable properties of polymeric
materials allow them to be used in unique situations where metallic and ceramic materials
fall short. Advanced technologies for individual use (such as conformal coatings for solar
power, which allow the user to coat solar material onto a surface of any shape [a feat
difficult to achieve for stiff silicon-based solar cells]) or kinetic generation of power from
flexible electronics in clothing or on skin could be used in everyday life and in survival
situations.
• Flexible electronics: These are polymer-based systems that exhibit all the properties of
standard electronics but can withstand bending and deformation. Exciting applications

18
exist in biomedical sensors, electronic skin, flexible televisions and phones, and soft
robotics.
• Extreme-environment energy generation: In regions where energy is harvested in
extreme environments—solar with little sun, wind turbines off-shore—instruments need
to be self-sufficient and able to withstand the constant strain of the environment.
Polymers with anti-fouling, water-resistant properties would improve delivery of energy
from ocean sites by preventing degradation of expensive underwater piping systems; self-
healing coatings capable of repairing inevitable damage to external components could
prevent the loss of solar cell function, usually caused by cracks on the cell surface.
• Enhancing efficiency of energy generation: Using polymers as active components in
energy generation is attractive for expanding the scope of the instruments. Incorporating
flexible, printable, transparent electronics such as organic thin-film transistors (OTFTs)
and organic photovoltaic (OPV) cells composed of polymeric semiconductors can
improve the performance of outdoor technologies and provide lower-cost alternatives
that surpass the current inorganic technology.[15]

1.4 Infrastructure, Communications, and Information


Polymers are already important in infrastructure, from the vinyl sidings on a house and
polymer pipes in plumbing to insulating foam and protective paints and coatings. How can the
next generation of polymers help here? New developments that enhance the lifetime of
infrastructure are key. Imagine paint that never chips or peels or pipes that don’t crack. New
developments using “smart”, environmentally-adaptive materials could lead to: paint that could
report damage to the infrastructure or be healed when scratched; sidings or coatings that are self-
cleaning or de-icing; insulating foam that can adapt heat transfer depending on environment; and
plumbing pipes that can self-sense damage and repair before a leak starts. These are just a few
considerations where new developments in polymeric materials and composites will have an
effect.
A different consideration is how the next generation of polymers can be used to affect human
sustainability. The current world population is ca. 7.2 billion and is projected to increase to 9.7
billion by 2050.[1] Combined with urbanization and recent mass migrations (due to war, famine,
drought, etc.), this increase is a huge challenge for human sustainability. Building the physical
infrastructure (housing, roads, communication networks, etc.) for this growing world is a
significant issue both from a societal and environmental perspective.
Grand Challenge: Rapid, low-cost, local production of lightweight, robust
infrastructure.
We are tasked with developing new integrated materials and smart technology that will
provide durable, healthy, environmentally conscious, aesthetically pleasing, lightweight, modular,
and inexpensive building materials to replace the current aging infrastructure and meet the needs
of a rising global population. The most common building material, concrete, is one of the largest
CO 2 producers, creating up to 5% worldwide of human-made emissions.[16] More than 70% of
oil/gas is burned to produce energy, so it generates CO 2 . Diverting these resources into producing
polymers could reduce infrastructural emissions by displacing concrete and also help meet the
growing demand for building materials. Polymers can address the growing need for increasingly
affordable, modular, durable, and flexible building/infrastructure materials. There are already
companies and groups across the world proposing to build houses[17] and roads (Figure 1.2) from

19
Figure 1.2. Computer desigwn renderings depict the proposed construction of a plastic road put forth by
VolkerWessels.

recycled waste plastic.[18] Such approaches allow prefabrication of the components and relatively
easy transport and construction (e.g., no need for cement to lock the polymer bricks in place).
Polymer production requires much less energy than metals, ceramics, concrete or, asphalt,
so it has a “green” advantage. As such there is a growing role for polymers in infrastructure (green
buildings, bridges, roads, coating/paints, etc.). Targeting polymeric materials specifically designed
for such applications offers tremendous opportunities that would simultaneously generate jobs and
stimulate the US economy.

1.4.1 Major Challenges/Opportunities:


In addition to large-scale availability, manufacturability, and cost, some technical aspects to
be considered include:
• Durability and restoration: New polymers that withstand extreme climate conditions
while retaining the physical robustness necessary to maintain infrastructural integrity will
be critical to future projects. Improved performance may be achieved by new innovations
in polymer composition and novel material architectures or morphologies.
• Fire resistance and safety: A common downfall of polymers used in infrastructure is
flammability and any resulting toxic degradation. Advances in flame retardation
technology have begun to affect industries such as insulation and fabrics, but continued
advances in research on flame resistant materials and coatings is necessary.
• Energy efficient light and thermal management: Energy efficient buildings are a constant
concern for home and business owners alike. Using polymers to improve thermal
insulation, create efficient light sources, and seal gaps in construction are just a few of
the ways to reduce energy waste in infrastructure.
• Integrated systems: By using advances in organic energy and electronic devices such as
optical fibers, organic thin-film transistors, and sensors (see Section 1.3), smart
technology can be embedded into new and existing infrastructure. Lightweight, flexible,
and robust polymeric systems can be used in applications such as paints and coatings
with sensing abilities; water pipes that can communicate flow rate and the presence of
contaminants; or roadways that relay traffic information and react to road conditions.
• Comfort and health: Developments in polymers for infrastructure must strive to improve
the overall comfort and health of the population. This includes finding ways to reduce
emissions through active construction materials, responsible manufacturing, and creating
ergonomic, aesthetically pleasing designs.

20
Communications present additional exciting possibilities for polymer science. As early as
the 1990s, polymer scientists noted a myriad of opportunities related to polymers in
communications and information technology, such as conducting polymers and optical/photonic
transmission.[19] For example, a key area for polymers is the development of fiber optic cables.
With their improved non-linear optical (NLO) properties compared to their inorganic counterparts,
polymer-based optical fibers can transmit information with more energy efficiency and minimize
energy use in server farms. Polymers can also affect the optical community through organic light
emitting diodes and holographic transmissions.
Although much progress has been made in this pursuit, advances portend exciting new
emerging technologies. For example, semi-conducting polymers may enable visible light
communication. This would enhance data transmission and help relieve the burden on the world’s
wireless communication network.[20] Controlled polymer architecture and internal phase behavior
and strategic manufacturing approaches have led to fiber optic composites with variable responses,
with potential applications in active monitoring and wearable sensors.[21] Encoding information
in macromolecules to read and write data is also an exciting next step.[22] Integration,
responsiveness, self-sensing, self-regulation, and self-repair are key features of polymeric
materials for these directions. Advances in communications also will include direct interfaces with
people, e.g., via interfacing with the brain. Lightweight, biocompatible, electrically-active
polymers may present enabling new technology.
Another class of polymeric systems related to communications is dielectric
polymers/polymer composites, which offer easier processing and flexibility than their inorganic
counterparts. One major use of dielectric materials is in the electronic industry. For example, a low
dielectric constant material can have faster switching speeds and minimize heat dissipation,
allowing more energy-efficient electronic devices. A key challenge for this class of polymer is
optimizing their use at the nanoscale and developing polymers/polymer composites that match or
surpass the dielectric properties of inorganic materials.

1.5 Water/Food/Air Applications


Large population growth and mass migrations present the world with widespread demands
for access to clean and safe water, food, and air. Currently, over 40% of the world’s population
lacks access to sufficient clean water, mainly through a lack of freshwater resources or lack of
infrastructure.[23] Increasing pollution from an ever-growing population creates issues of water
and air purity and safety. Recent examples include the air purity problems plaguing highly-
populated areas in China and India. Food availability and safe storage are also global concerns.
Polymers will play an integral role in solutions to these problems. For example, major
developments are needed to dramatically improve access to water and food (e.g., improved
membranes for water purification/desalination; barrier materials for better food preservation).
Water and air pollution abatement (e.g., polymer membranes) will be a big consideration as the
worldwide population grows.
Grand Challenge: Combined 100x improvement (cost, efficiency) in new purification
materials and the development of biodegradable, smart, high-barrier
polymers for food packaging.

1.5.1 Major Challenges and Opportunities


We highlight a few of society’s pressing demands that will be met by innovative solutions
enabled by research and development on polymeric materials.

21
1.5.1.1 Water
Clean water should be considered a basic human right and as such a global goal is access to
affordable, energy-efficient, sustainable water for all. To accomplish this, there are a number of
major challenges ahead, many of which can be addressed by research in the polymer field.[24]
Specific topics include water purification using polymer membranes (Figure 1.3),[25] detecting
contaminants and remediating water using polymers as sensors, and employing polymeric
materials for efficient, long-distance water transport.

Figure 1.3. Having access to clean water across the world is a key challenge for the next decade. One possible
way to achieve this is with next-generation polymer membranes.

Advances in membrane technology for water purification are critically reliant on polymer-
based innovations. The use of pressure-driven membrane filtration to purify water from diverse
sources (e.g., seawater or industrial wastewater) is a cost-effective way to address the emerging
problem of water sustainability. However, there are still major technical barriers to be overcome
to improve the energy efficiency of filtration. Critical parameters to consider in all membranes,
polymeric or otherwise, are permeate flux, efficiency/selectivity of contaminant rejection, and
conditions required for purification (e.g., do they need high pressure or temperature?). Key goals
for future filtration processes are: membranes that achieve high flux, high contaminant rejection
and work under low-cost, energy-conservative conditions (atmospheric temperature and pressure).
Gravity-driven and solar-driven filtration fits this definition and would require less infrastructure
than pressure-driven processes, but the flux of current filters is not high enough for large-scale
water purification. Reducing the energy requirement of pressure-driven membranes, i.e., higher
flux while maintaining high contaminant rejection at lower pressures, is an important challenge.
Can these goals be met with new approaches/designs that can be imagined for the next
generation of membranes? One approach is the concept of incorporating “directed water channels”
in the barrier layer. Membrane permeability would be improved based on the mechanism of size
exclusion, inspired by the presence of water channels formed by aquaporin proteins. Designing
polymer membranes with inherent anti-fouling and anti-scaling properties would prevent the
buildup of contaminants over time, increase flux over the life of the membrane, and eliminate the
need for environmentally-toxic disinfectants or anti-scaling agents. One of the major challenges in
accessing clean, fresh water is desalination of seawater. Current membrane methods using reverse
osmosis (RO) already have the advantage of greatly reduced energy requirements compared with
distillation.

22
Beyond simple filtration, polymer membranes will be designed as smart materials that purify
water and monitor the presence of contaminants. Applying stimuli-responsive configurations
within a polymer membrane would let the membrane change color, shape, or conductivity based
on its programmed response to key chemicals in the water. Alternatively, membranes could be a
way to recover selected contaminants, possibly by attraction and trapping mechanisms rather than
simply blocking them. This offers the potential of purifying the water plus the advantage of
recycling rare or limited access materials, providing additional economic advantages. A wide
range of sequestered targets can be imagined including metals (e.g., gold, lithium, platinum,
palladium), phosphorous-containing and nitrogen-containing compounds (e.g., from agricultural
run-off), and even some organic compounds, such as pharmaceuticals.
In addition to membranes, polymers can play expanded roles in other aspects of water
purification, such as acting as adsorbents, flocculants, and antimicrobials; and in photocatalysis, a
pretreatment used to remove pathogens and trace contaminants.

1.5.1.2 Air
Like clean water, access to clean air should be considered a basic human right. Air pollution
is a global problem, if only because of climate change. Air pollution can come from many sources,
including changes in climate (e.g., increase in wildfires during drought), natural sources (e.g.,
volcanos), and increases in population densities. However, most of the air pollution comes directly
from human activity, which includes transportation sources, stationary sources (industrial
facilities, power plants, solid waste dumps, etc.), area sources (urban fireplaces), and agricultural
sources (e.g., livestock, fertilizers). As such, an increasing population comes with increasing air
pollution. The human (mis)usage of land can also lead to significant air pollution (e.g., the Dust
Bowl during the 1930s on US and Canadian prairies).
Polymers are crucial to confronting air pollution, including use as purification membranes
or as sequestering agents for pollutants. Polymer fibers are already a primary air filtration method
for personal use, such as respirators, car filters, and air conditioners (Figure 1.4). Accessing
membranes with high flux and high selectivity is an urgent goal. Major air chemical pollutants
include CO 2 , acid gases (SO 2 , NO X ), halogenated
hydrocarbons, ozone, methane, and solid
particulates. Solid particulates are a distinct
concern, as they can penetrate deep into the lungs
and bloodstream and have been linked to myriad
diseases, including cancer, asthma, and respiratory
failure. They are classified as PM 10 or PM 2.5 ,
meaning particulate matter of less than 10 or 2.5
micrometers in diameter, respectively. More
efficient and economical polymer fiber membrane
technologies are needed to meet these demands.
Removing CO 2 from flue gas is critical to Figure 1.4. According to the World Health
Organization one in eight deaths is linked to air
reducing this greenhouse gas in the environment. pollution.[26]
The annual amount of CO 2 emissions from
burning fossil fuels is ~35 billion metric tons
worldwide. Polymer membranes with high selectivity for CO 2 (ideally should remove > 90% of
CO 2 from flue gas), high throughput (flux), good mechanical and chemical robustness, and low
cost will have a major effect.

23
1.5.1.3 Food
Polymer science has contributed in major ways to feeding the world population. New
packaging technologies have revolutionized how we store, preserve, transport, and market meats
and produce, dramatically extending the life of food and significantly reducing waste. For
example, vacuum packaging of beef in an oxygen barrier film extends its shelf-life from 4 days to
up to 30 days.[27] Polymer packaging is lighter compared to glass or metal, which makes
transportation cheaper. A primary focus is improvement of the selectivity of polymer barrier
materials for packaging. Designing and developing polymeric materials that can selectively
regulate the diffusion of moisture or oxygen and uni-directional barrier materials that allow gases
to flow in one direction but prevent the reverse flow are two major challenges.
Another emerging technology in the plastic packaging industry is active or smart packaging.
Using either properties inherent to the polymer itself or imbibing an existing polymer with
molecular or electronic sensors, scientists are able to dramatically enhance the use of the packaging
itself by generating materials that go beyond the typical seal-and-store application. For example,
consider packaging food in a plastic wrap that turns a color when the perishable product has gone
bad or a film containing anti-bacterial or scavenging properties that significantly improve the shelf-
life of food (Figure 1.5). The industry is plagued with challenges to deliver fresher, healthier food
to the general population. Beyond the implementation of improved overall barrier properties, the
ability to maintain a consistent barrier under extreme conditions—deserts, oceans, extreme cold—
is an important factor. Polymers are uniquely qualified to handle these demands offering access to
a wide range of mechanical properties. This is demonstrated by some of the most prevalent
commodity polymers, such as polyethylene, which makes up both our plastic shopping bags and
much of our indoor water plumbing.
Advancements in processing commercial and developmental polymers are a major area of
focus for the food packaging industry. One
of the great advantages of plastics and
elastomers is that these products can be
processed into almost any shape, which
allows them to be economically fabricated
into water bottles, clear plastic wrap,
insulating foam carry-out containers, and
freezer bags. Continued improvement in
food packaging will directly affect the
global economy. Developments such as
multi-layer films for simultaneous UV
protection and oxygen barriers will affect
the shelf-life and quality of all types of
food. Integration of computational
Figure 1.5. An example of active packaging design for the methods, such as finite element analysis,
increased lifetime of perishable items. A food packaging into the processing community has
company (Paper Pak Industries) demonstrates their patented revolutionized the way industrial
design for the UltraZapXtendaPak, featuring a CO2- processing facilities evaluate polymer
producing pad that keeps the food fresher, longer.[28]
manufacturing methods and operations.
Continued growth of these techniques is
crucial to development and to implementation of the latest advances into the commercial market.

24
Polymer science and engineering also has an enormous effect on farming and agriculture.
Polymers absorbents retain water in the ground and thus reduce water requirements; remove
contaminants and toxic chemicals to remediate soil; control release of biocides, herbicides, and
pesticides; and as a barrier film, protect crops. It is hard to imagine how a modern farm could
function without plastics and elastomers, and the importance of such products will continue to
grow.

1.6 Health
“Biomaterials already play central roles in the $200-billion-per-year medical device
industry, and they currently improve the quality of life for millions of people
throughout the world. Further advances in this industry will be critically dependent
upon fundamental studies of interfacial phenomena, cell-material interactions, the
synthesis and characterization of particulate systems, and the development of
materials for use in sensors and diagnostics. An ambitious vision for the future of
biomaterials research might include the creation of cell-powered medical implants,
virtual patients, materials that anticipate and prevent disease, and implanted devices
that restore lost function and adapt and grow with the patient. Realizing this vision
will require investment in the kind of fundamental research that lies at the core of the
NSF mission.”—2012 NSF Biomaterials Workshop.[29]

The field of biomedical materials is growing rapidly, enabled largely by advances in


polymeric materials. Examples range from plastic bags to store and preserve biological fluids (e.g.,
blood bags) to replacement hips, knees, heart valves. A 2012 NSF Biomaterials Workshop and
corresponding report focused on all aspects of biomaterials, including sections on (1) Hard
Materials and Composites, (2) Soft Materials, (3) Cell-Materials Interactions, (4) Dispersed
Systems and (5) Thin Films and Interfaces.[29] There is significant overlap between that report
and the advances in polymers anticipated by this report. The goal of this report is to complement
rather than reproduce the previous report by focusing on recent trends and emerging needs in the
field of biomedical polymers. One trend in Societal Needs as it relates to human health is the
growth in human infectious diseases. From the Ebola outbreak in West Africa in 2014 to the Zika
virus outbreak across the Americas in 2016 (Figure 1.6), infectious diseases are constantly in the
news. While non-infectious diseases still represent the leading cause of mortality in most
developed nations, a recent study[30] showed there has been a dramatic increase in infectious
disease outbreaks in the last 35 years. Furthermore, zoonotic diseases (those which can be
transmitted to humans from animals), a class which includes Ebola, HIV, bubonic plague, and
Lyme disease, also showed a significant increase in rate of incidence over the last few decades.
Grand Challenge: Innovative materials development for cost-effective infectious disease
control, treatment, and diagnostics.

1.6.1 Major Challenges and Opportunities


There are a number of health-related major challenges where polymers can play a role in the
future. A few of the key major challenges identified during this workshop are outlined here.
New materials development for infectious disease control. Materials Development for
Infectious Disease is an area that recognizes the growing concerns of drug-resistant infectious

25
agents (MRSA, VRE, etc.), and the
increased risk of wide-spread infection
with human mass migrations. Generally,
each bacterial infection is composed of not
one micro-organism but thousands of
aerobic and anaerobic bacteria and fungi.
The human body depends on myriad
complex mutualistic and commensalistic
relationships with symbiotic
microorganisms (the microbiome) to
remain healthy. A major challenge in
treating disease is to selectively target
pathogenic microorganisms rather than Figure 1.6. Zika virus outbreak has been linked to the
increase in microcephaly in children born to women
beneficial ones. Innovative polymers to infected with the disease while pregnant.[31]
prevent infection and treat infection need to
be developed. These polymers should be designed for specificity towards pathogens, economy in
cost, and stability over a broad temperature range (-20 °C to 140 °C) to allow broad distribution.
These polymers, for instance, could be used to treat surfaces to prevent infection and to treat
infection in a site-specific (localized), affordable, rapidly-deployable manner.
Overcoming physiological barriers to treat disease. Several diseases and disorders that
critically affect our society have been very difficult to treat, primarily due to the challenges of
targeting drugs and therapeutics across hard to cross physiological barriers. Crossing the blood-
brain barrier is a serious challenge that has made the treatment of neurological disorders and
diseases (e.g., brain cancer and meningitis) particularly difficult. The ability to transport molecular
inhibitors, biologic drugs, siRNA, mRNA, or DNA across this barrier may open up new therapies
for chronic disorders such as Alzheimer’s and Parkinson’s. A key promise of the Brain Initiative
is to gain better access and understanding of the brain, and polymeric carriers and imaging agents
may also help leads to development of disease treatment. Other key physiological barriers include:
endothelial tissues in the lung, which can provide more efficacious and targeted treatments of
infectious disease and chronic respiratory disorders; and epithelial barriers found in skin and in the
stromal tissues of tumors that can assist in addressing cancer, developing new vaccines, and
accessing the body’s immune system. Addressing targeted delivery to the gut involves traversing
mucosal and cellular barriers. Polymeric systems, including polymeric nanoparticles, offer a
unique advantage in traversing these barriers. Access to a range of synthetic compositions, tunable
charge, hydrogen bonding, hydration, and conjugating biologically relevant moieties and cues are
all feasible. The range of mechanical properties polymers present, from rigid to soft and pliable,
the potential to vary these properties with external stimuli, and the range of sizes and shapes that
polymer systems can provide further enable access across physiological barriers.
Regeneration of tissues. Regeneration of tissues (soft and hard) is a highly-recognized need
worldwide. One area of soft tissue regeneration is that of skin (e.g., in the case of burns or cuts).
Burn wounds heal through a repair mechanism, creating scars that cause tissue contracture.
Contracture can be painful and debilitating on the face and other flex points (e.g., joints). Scar
contracture can result in inability to chew, see, effectively grasp objects—all of which can create
a sense of isolation.
With our population aging and obese, millions of people worldwide suffer from debilitating
chronic wounds, such as diabetic ulcers and pressure ulcers. Under current standard of care, these

26
wounds routinely require several months to several years to heal. When they heal, the dermis is
repaired/damaged, not regenerated, which increases the likelihood of re-injury. In cases of acute
bodily injury, bone repairs often become impossible due to detached bone fragments or gaps in the
injury that are too large for the body to heal without help. Bone tissue regeneration, especially in
load-bearing sites, requires a strong yet moldable material that can bridge gaps in the bone to
promote new cell growth and simultaneously support the injured area. Regeneration of tissues
rather than repair would allow the injury to heal with nascent-like tissue. It would require materials
with bio-mimicking, smart properties, at least. Polymers, for instance, can be a barrier to
exogenous contamination, a gas/ion transport membrane, a substrate for cell
migration/proliferation, or a delivery device. Broader effects and fundamental understanding of
artificial and biological tissues are needed. Another key area is developing improved biomedical,
polymeric adhesives for injuries and surgical incision sites.
Rapid affordable diagnostics. The rising costs of healthcare in both industrial and
developing nations, combined with the ever-present need for a more uniform global distribution
of diagnostic methods has driven research toward the development of diagnostic approaches. For
example, lightweight, portable, and easy to use approaches to detecting disease can enable a much
more rapid response time in the case of disease outbreaks. The ability to detect glucose levels for
diabetics using cheap disposable sensors, or to determine important metrics such as cholesterol
level and fat content, along with modes of remote monitoring of patients with ease, could lead to
disruptive approaches in addressing health and decreasing the costs of the global health care
system. Polymers provide the ability to embed function, responsive properties, and highly selective
detection capabilities and to integrate them into paper or plastic substrates or devices. Furthermore,
polymers can be designed to stabilize sensing components, enabling ease of shipping and the
elimination of cold chain storage.
Minimization of animal models for toxicity and biocompatibility testing. Historically, using
animals to determine human safety and efficacy has been the gold standard. These pre-clinical
models require lengthy times to determine response; heavy upfront costs for a YES/NO answer;
and the loss of animal life. Using polymers, polymer matrices, 3D scaffolds, etc., to mimic
biological system responses can address the concerns and minimize or replace animal testing. For
example, when drugs or vaccines are transported without adequate packaging to hot or humid
climates, the benefits of the compounds are lost on the target population. As packaging materials
improve, testing for low thresholds of gas permeation through them becomes a critical component
for evaluation without animals.

1.7 Aspirational, Inspirational, and Curiosity-Driven Needs


“[L]et us try to imagine a dweller in the “Plastic Age” that is already upon us…It is
a world free from moth and rust and full of color, a world largely built up of synthetic
materials made from the most universally distributed substances…a world in which
man, like a magician, makes what he wants for almost every need, out of what is
beneath him and around him…”[32]

The passage above describes Yarsley and Couzens’s vision of plastics, 75 years ago, in 1941.
Since then, plastics have indeed infiltrated nearly all aspects of modern life, from commodity to
specialty products. While the concept of societal needs certainly encompasses the societal
challenges detailed above, not all societal needs are reactive—trying to alleviate or solve societal

27
problems. A key interface between science and society is how scientific advancements can relate
to and fulfill society’s aspirational, inspirational, and curiosity-driven needs.
Consider for a moment the complexity of the functions of a single cell. While the science
has long been aware of the existence of cells and has extensively catalogued the functions of many
of its parts, the level of compartmentalized sophistication within the cell has yet to be synthesized.
Polymer scientists, with their knowledge of macromolecules and macromolecular behavior, are
uniquely positioned to study and predict the biomechanical processes within a cell and design a
way to create a synthetic cell. The implications of synthetically-designed cells would be profound,
but the level of knowledge needed to achieve this lofty goal can only be attained through significant
curiosity-driven research.
The ability to more effectively interface biology with synthetic materials has a lot of
potential. For example, interfacing conducting polymers (those that can conduct electronic,
molecular, and ionic bio-signals) with the body, namely organic bioelectronics,[33] offers
opportunities to regulate biological processes or to sense and monitor them. One specific area of
interest is in developing better bioelectronic interfaces with the brain, which could treat diseases
affected by electrical simulation e.g., Parkinson’s.
Space exploration is another shining example of where aspirational polymer science and
engineering research has played a role. Without the exceptional material properties required for
humans and technology to survive in space, our world would be devoid of such inventions as
modern insulation, memory foam, and scratch resistant lenses—all made from polymers. As
science reaches deeper into space, highly-engineered, lightweight materials are needed to meet the
demands of extreme environments and, for humans, long-term living in confined, resource-limited
spaces. These innovations can lead the way to solutions for earth-bound problems—advances in
computer technology for long space missions could translate into driver-less cars; lighter, more
robust plastics for spacecraft interiors and insulation could decrease the price of transportation;
closed-system recycling of water and food during human missions might lead to breakthroughs in
water purification and distribution in remote regions.
Of course, some of this can be more directly application-driven. For instance, extraordinary
advances in electric energy storage and artificial intelligence use have driven the dream of personal
flying vehicles closer to reality. Or consider how new adaptive/protective materials have enabled
stronger, lighter, and superhuman-like armor to outfit soldiers in combat. Often these new frontiers
are introduced through “curiosity-driven” research, i.e., research without a targeted, “real-world”
application. It is such research that can bring disruptive technologies to society. For polymers,
many historical examples of accidents being translated into serendipity support this trend, from
spinning the first synthetic nylon fibers to manufacturing Teflon™ coatings. Society needs to
support this type of fundamental, long-term research; the US needs to maintain leadership in
polymer science and engineering worldwide. Exciting opportunities that extend beyond the
traditional realm of polymer science and engineering, such as brain-medicine communications,
soft membranes for robotic sense of touch, and flexible electronics await innovations from
inspirational aspirations.
Grand Challenge: Invest in an environment that allows and encourages scientists,
engineers, and entrepreneurs to pursue bold, transformative ideas
and think outside the box.

28
1.8 Inclusive Societal Outreach and Education
To maintain global competitiveness in technological advancements, greater inclusion of
individuals with different backgrounds and experiences is of utmost importance. Figure 1.7 shows
the historic and projected demographics of the US population over a century since 1965.[34] The
trends indicate a changing, more diverse society ahead.

Figure 1.7. The evolving US population, 1965-2065. Select years are highlighted for reference. White, Black,
and Asian include only single-race non-Hispanics. Asian does include Pacific Islanders. Hispanics are of any
race. Projections of U.S. population are based on assumptions in birth, death, and immigration levels.[34].

Grand Challenge: Foster recognition and support for diverse groups of interdisciplinary,
educated scientists and engineers alongside a technology-literate
society.
To increase diversity in the science and engineering workforce and bolster scientific literacy
within society, more inclusive outreach and education need to be developed. This should take place
across educational institutions and be embraced by citizens and communities nationwide.
Programs should emphasize and foster curiosity and problem-solving skills at younger ages. The
programs should be made widely accessible for children and their families, inside and outside the
classroom. Such mentorship efforts should not focus only on underrepresented ethnic groups but
also encourage women to pursue careers in science and engineering. These efforts should include
other underrepresented communities such as the visually-impaired or the deaf/hard of hearing.
Educating a diverse pool of talented young people will support and guide unrealized potential and
dreams into tangible discovery and self-fulfillment. In addition, solving the grand challenges
outlined above and helping define the next set of grand challenges will require polymer scientists
and engineers with diverse backgrounds; with experiences that bring different perspectives to
problems; who can increase public visibility; who can advocate for science and engineering
education across the whole of society. Communication from an increasingly diverse group of
leaders in these fields can bridge various disconnects and misconceptions. Altogether, broader
efforts towards reaching and educating all individuals will cultivate new ideas, help solve

29
imminent societal needs, and strengthen the future US workforce. Sections 7 and 8 of this report
outline specific ideas and actionable recommendations to address major challenges and
opportunities in education, diversity, and outreach.

30
SECTION 2: MACROMOLECULAR SYNTHESIS
Discussion Leader
Geoffrey W. Coates, Cornell University

Group Members
Anna C. Balazs, University of Pittsburgh
Lisa S. Baugh, ExxonMobil
Padma Gopalan, University of Wisconsin—Madison,
Robert B. Grubbs, Stony Brook University
Marc A. Hillmyer, University of Minnesota
Krzysztof Matyjaszewski, Carnegie Mellon University,
Thomas F. Miller III, California Institute of Technology
James C. Stevens, The Dow Chemical Company
Brent S. Sumerlin, University of Florida

Special Acknowledgements
Kathryn L. Beers, National Institute of Standards and Technology
James M. Eagan, Cornell University

2.1 Introduction
Grand Challenge: Macromolecular syntheses afford the opportunity to create an infinite
spectrum of tailored molecular architectures.
Realizing this potential necessitates reliable, sustainable, and cost-effective methods to
synthesize macromolecules of any primary structure and architecture, informed by the quest for
discovery and societal need. The recommendations in this section of the report are designed to
meet the grand challenges of macromolecular synthesis. By overcoming these barriers, polymer
science and engineering can expand and improve the properties, applications, and understanding
of polymer materials.

2.2 Significance, current state-of-the-art, and challenges


2.2.1 To synthesize macromolecules of any primary structure with reasonable time frame
and cost.
The last decade has witnessed significant progress in the synthesis of macromolecules with
precisely controlled architecture. Generally, well-defined polymers are prepared by
controlled/living polymerizations (CLPs), with concurrent growth of all polymer chains (fast
initiation) and a diminished contribution from chain breaking reactions (transfer and termination).
In this manner, polymers with low dispersity, approaching Poisson distribution (Mw/Mn ~ 1), are
prepared.
Originally, living polymerizations were developed for anionic polymerization and
subsequently expanded to ionic ring-opening polymerization, coordination polymerization of both

31
olefins and cycloolefins (ROMP), and eventually to radical polymerization. Classical living
polymerization requires all growing polymer chain ends to be mediated by controlled agents that
can be relatively expensive, e.g., metallocenes, post-metallocenes, or stable free radicals. But
recently, various catalysts that can be used at the parts per million level have been developed for
use in the presence of less expensive reagents, such as alkyl halides in atom transfer radical
polymerization (ATRP) and alkylzinc or alkylaluminum compounds in degenerative transfer
shuffling catalyst systems in olefin polymerization. This approach very significantly reduced the
cost of commercial synthesis of various block copolymers. A challenge is further reduction of the
cost and development of more environmentally friendly conditions in these processes.
Several elements of macromolecular architecture can be controlled in CLPs. They include
chain topology, chain composition, chain functionality, chain stereo-structure, and chain
uniformity. These elements are based on chains with covalent bonds connecting monomeric units.
In addition, dynamic non-covalent bonds can also form macromolecular chains with properties
strongly affected by the dynamics of chain interactions (dynamers, and vitrimers).[1]
Chain topology elements span from linear chains to cycles and various branching features.
They can include long or short chain branching, loose or dense branching (or even hyperbranched
systems), and eventually dendrimers. Branches can be distributed with a different density along
the chain (as in graft copolymers) or very densely (as in bottlebrush copolymers). Branches can be
limited to one focal point (as in star polymers which can be formed by arm-first or core-first
approach). They can also be distributed in the bulk material, forming networks with different
uniformity, mesh size, and composition. The type and degree of branching can tremendously affect
mechanical or rheological properties of the resulting polymers. A challenge is how to design
branching degree, uniformity, or location; correlate these elements with macroscopic properties;
then precisely carry out synthesis of such materials.
Functional groups can be placed in various parts of macromolecules. They can be located
with a certain density along the polymer backbones, at the extreme position of chains, including
chain ends in telechelics, chain center, ends of arms in stars and bottlebrushes or in the cores of
stars, or chain ends for hyperbranched or dendritic molecules. These groups should carry specific
functions that can be used for further reactions, crosslinking, or attachments of other moieties—
whether it is a biomolecule, a drug, an optoelectronic material, or another species. A challenge is
to incorporate moieties in a specific position of macromolecules with reactive orthogonal
functionalities for further reactions and synergistic effects.
Control of stereostructure is essentially limited to coordination polymerization of both vinyl
and cyclic monomers, with some limited control in anionic systems. Specially designed catalysts
can provide excellent and programmable stereocontrol for polypropylene and other polyolefins
and can be even extended to stereoblock copolymers. A challenge is to take such control in radical
polymerization, improve it, and extend it to (meth)acrylates, acrylonitrile or styrenes.
CLPs can provide polymers with very low dispersity, essentially a Poisson molecular weight
distribution (MWD). However, there is a growing interest in synthesis of materials with higher
dispersity and even an intentionally programmed MWD. This can be accomplished by various
exchange reactions. They can include depropagation, transesterification, or transacetalization
reactions but also a designed slow exchange between active and dormant species, an exchange
slower than propagation. This can be accomplished by using a very low concentration of Cu
catalysts in ATRP or by using less efficient RAFT reagents (dithiocarbamate for methacrylates).
MWD might also be controlled by slow feeding of initiators or by a continuous termination
process. Programmed feeding (constant rate; linear or nonlinear addition rate) can change

32
skewness and asymmetry of MWD (Figure 2.1).[2] Block copolymers with broader MWD may
have lower ODT; form larger domains; generate new bicontinuous morphologies; show improved
performance in a number of applications, including pressure sensitive adhesives; and even exhibit
photonic properties. Thus, controlled and designed heterogeneity, including MWD, tacticity,
composition (gradients), and shape (branching) can lead to a new class of materials with novel
properties.[3] It is a challenge to harness these parameters and prepare advanced materials with
broader processing windows, broader glass transition, improved fracture toughness, higher
critical micellar concentrations, and new morphological features.
CLPs are typically carried out under
Initiator
Relative constant temperature or pressure.
Time
of However, it is advantageous to use several
Monomer Chain
Initiation
external triggers to control polymerization
Controlled MWD rates and loci. Such temporal and spatial
Tuned Polymer
Composition
control can be accomplished by light,
Figure 2.1. Controlled and continuous addition of initiator electrical current, mechanical forces, or
allows tuning of the molecular weight distribution during
living chain-growth polymerizations. This extends the
chemical input (pH). The remaining
control over targeted molecular weight, which has become challenge is how to use them together and
the hallmark of living polymerization techniques. Such provide dual control, start and halt
techniques also include control over molecular weight polymerization at will, and perhaps switch
distribution composition (adapted from [2]). from one mechanism of polymerization to
another, expanding range of accessible
elements of polymer architectures.
New CLPs, especially proceeding by radical means, open new avenues to prepare hybrid
materials. They include organic/inorganic hybrids based on nanoparticles, nanotubes and flat
surfaces; but also include bioconjugates formed by covalently linking natural products, proteins,
nucleic acids, and carbohydrates, with synthetic polymers. Covalent grafting of polymers from
inorganic surfaces provides access to very densely grafted brushes that affect many properties,
including enhanced lubrication, antifouling, or antimicrobial. Nanoparticles, nanorods, or
nanotubes with densely grafted polymers do not aggregate, and they disperse very well in solvents
or polymer matrices. Proteins with grafted polymer chains can circulate for a longer time in the
human body, can survive at low pH, can be dispersed in organic solvents, and be used as catalysts
at higher temperatures. Nucleic acids combined with polymers can self-assemble, pass efficiently
through cell membranes, and form various polyplexes. The challenge is how to design the most
efficient materials and how to carry out their synthesis precisely.
Precision polymers have been prepared by chain growth polymerization of vinyl or cyclic
monomers. By contrast, conjugated polymers (comprised of alternating -bonds) are
predominantly synthesized by a step-growth process using metal-catalyzed cross-coupling
strategies, with limited control over molecular weight, dispersity, and architecture (Figure 2.2).
The development of chain-growth polycondensation for conjugated aromatic polymers, mostly
thiophene derivatives, via catalyst-transfer polycondensation provides access to well-defined
materials with low dispersities, controllable molecular weights, and predetermined sequences.[4]
There are many opportunities to explore more complex -systems via controlled polymerization.
Improved control and development of a wide variety of coupling procedures are needed to link the
aromatic units into long chains. Advances in metal-ligand design are necessary to enhance the
mechanistic understanding of the binding event between the metal and growing polymer
fragments; and to design more complex and functional conjugated scaffolds. The bulk properties

33
(mechanical, electrical, optical, or thermal) and the solid-state organization of the materials should
be significantly better tuned through control of backbone composition and shape. It is a challenge
to prepare well-defined conjugated polymers via a chain growth from the large number of already-
explored conjugated building blocks.

Structure and Organization


Enhanced Understanding of Metal-Binding
M n control / low MWD Block Copolymers
ML n

Br Het Het Het Br Graft Stars


Metal-ligand design strategies
Diverse classes of building blocks
Improved control in coupling procedures

Figure 2.2. Well-defined conjugated polymers via chain-growth catalyst transfer polymerization.

2.2.2 To polymerize renewable monomers under the principles of ‘green chemistry’; to


create cost-competitive high-performance materials that are environmentally benign;
or to create materials that simply last forever (durable).
Historically, the goal of polymer synthesis was often to prepare macromolecular materials
with a certain set of physical, chemical, or material properties. Polymeric materials have played
such important roles in technologies directly benefiting society that most attention was given to
achieving the desired material. No one asked: where does the starting material come from? What
catalysts will be used to prepare the polymer? Is the polymer safe in use from a human health
perspective? At the materials end of life, how is it disposed? What are the environmental costs of
making, using, and discarding these materials? Such questions were simply not at the forefront for
the synthetic polymer chemist. This situation has changed, and sustainability is now an ever-
present concern across the globe.
Renewable resources. Nearly all polymers come from petrochemical resources, such as oil
and natural gas. While transportation fuels dominate their use, it has been estimated that upwards
of 20% of these resources will go to polymers production by 2050. There is one principal reason
for using these finite resources for polymer synthesis: they are inexpensive. To shift to annually-
renewable plant-based starting materials, converting biomass to established monomers (like
ethylene and propylene) and to new-to-the-world monomers (like lactide) must be improved. The
grand challenge for the polymer synthesis community is to partner with metabolic engineers,
organic chemists, catalysis experts, and chemical engineers to discover and develop new strategies
that contribute to the efficient and low-cost conversion of abundant (and non-nutritive) biomass
into monomers for polymerization. Important recent work in this arena includes the catalytic
conversion of sugars to lactic acid using zeolites,[5] the preparation of terephthalic acid from
cellulose derived molecules,[6] and fermentative processes to give isoprene.[7] In addition to
abundant biobased resources, using carbon dioxide for polymer synthesis is also very attractive
given its prevalence and the new developments in the catalytic conversion to polymers with
interesting and attractive properties.[8]
Green (polymer) chemistry. The development of polymerizations that can be done at room
temperature and low pressure to reduce energy costs; the use of non-toxic metals for the catalytic

34
conversion of monomers to polymers; and the use of safe (or no) solvents in polymerizations
should all contribute to more environmentally-friendly polymer synthesis protocols. Monomer,
polymer, and polymer additive toxicity are other elements of concern during and after polymer
synthesis. The community should be aware of these issues and address toxicity as early in the
process as possible.
Designed degradation. One of the key features of polymeric materials is their durability. Of
course, this is exactly the property that leads to their persistence in the environment. A grand
challenge for polymer synthesis is the preparation of materials that are durable in use but that can
degrade to innocuous byproducts on demand or upon the action of an external stimulus. Achieving
both durability and degradation has huge challenges, and clever approaches have been explored.
However, the ability to have the best of both worlds has eluded us, and new approaches should be
considered.
Chemical recycling. Polymer recycling suffers from two key problems. First, the process of
mechanical recycling and reprocessing often results in inferior properties, compared to the virgin
material. This is often due to some chemical or molar mass degradation that has compromised
mechanical properties. Second, mixing recycling streams leads to polymer mixtures that can
dramatically reduce the quality of the recycled materials. While there have been some successes
in compatibilizer development, this area is far from mature and warrants more attention. Chemical
recycling back to the original monomers or to new products that have value is a very important
and emerging area of research with outstanding potential. Instead of “downcycling” polymeric
materials, the community should emphasize the development of chemical recycling that efficiently
uses waste polymers as a valuable resource
for other products. Recently, there has been
very promising activity in this area with
aliphatic polyesters (Figure 2.3),[9]
chemical recycling of carbonates,[10] and
even with the recalcitrant
polyethylene.[11] This is a fertile area, and
polymer chemists should continue to think
about polymeric materials as feedstocks for
other useful chemicals or products that can
have value in other technologies.
Figure 2.3. Chemical recycling of poly(butryolactone) Repurposing thermosets.
using an organocatalyst.[9] Thermosetting polymers have the distinct
advantage of being chemically and
thermally resistant materials used in many high-performance applications. They generally cannot
be reprocessed and thus recycled and reused. With the recent advent of vitrimers, this is changing.
Vitrimers are reprocessible thermosets that derive their properties from intermolecular associative
exchange processes (Figure 2.4). This new class of materials has revolutionized the thermoset
world and holds tremendous potential.[12] The challenges in this area are developing facile
exchange processes, broadening the scope of materials that can benefit from this associative
exchange process, and achieving the ultimate high performance enjoyed by traditional thermosets.
Polymer chemists at work on this problem must focus on polymer-polymer reactions instead of
the more typical monomer to polymer reactions; this will require thinking about polymers as
reagents. Thus, designing backbones capable of associative exchange will help develop a broader
range of vitrimeric materials.

35
Figure 2.4. Vitrimers based on polyesters as reprocessible thermosets.[12]

2.2.3 To develop a theoretical framework for screening polymer targets that is (i) reliable,
(ii) atomistic-scale in resolution, and (iii) efficient enough to provide answers rapidly.
Computational methods offer extraordinary potential to aid discovery, characterization, and
synthesis of polymer materials. Deliverables range from fundamental mechanistic insights about
polymer interactions and transport properties to prediction of new equilibrium and non-equilibrium
polymer phases to rank-ordered screening of polymers for a property of interest. However, the
simulation and modeling of polymer materials is complicated by technical challenges, including
the extraordinary slow relaxation times of polymers compared to local molecular motions and the
subtle, many-body character of non-bonding (Coulomb and dispersion) interactions among
polymer chains.
A grand-challenge problem for advancing polymer synthesis is thus developing theoretical
frameworks for computational modeling of polymers that are (i) sufficiently accurate and reliable
to describe the properties of interest, (ii) atomistic-scale in resolution to account for detailed
interactions and the effects of subtle structural changes, and (iii) efficient enough to explore huge
swathes of chemical space and to provide answers rapidly, thus enabling useful iteration with
experimental collaborators (Figure 2.5).
The current landscape for the
simulation and modeling of candidate
polymers. At the atomic scale, the state-of-
the-art for describing polymer interactions
with high accuracy is given by ab initio
molecular dynamics (MD) simulations that
use a first-principles density functional
theory (DFT) electronic structure
description.[13] However, even with
massively parallel implementations, this
approach is limited to small system sizes (<
500 atoms) and short timescales (< 10 ns),
which are insufficient to characterize the
long-timescale dynamics of polymers. At
Figure 2.5. The screening and discovery of promising the other extreme, empirical force fields
polymer materials presents extraordinary challenges for have been used in many MD simulations of
molecular simulation methods.
polymers, such as polarizable force-field
36
methods,[14] the multi-state empirical valence bond method,[15] symmetry-adapted perturbation
theory,[16] and embedding methods.[17] The robust, general, and efficient description of polymer-
polymer, polymer-ion, and polymer-molecule interactions at the atomic scale remains an important
and outstanding challenge for theoretical methods development.
A variety of coarse-grained and multi-scale models have been developed to enable
simulations that approach experimentally-relevant timescales for polymer dynamics. The most
straightforward and widely-used strategy reduces the atomic resolution by combining individual
atoms into collective “beads” or “super-atoms” that interact via pairwise potentials.[18] This
coarse-grained approach can reduce the total number of pairwise interactions to be computed and
increase the characteristic timestep with which the system can be simulated using MD. It has been
used successfully to characterize many aspects of homo-polymer and block-copolymer phase
behavior and rheology,[19, 20] although it is somewhat limited to the description of long-length
scale features of polymers that are governed by the non-specific interactions of the coarse-grained
beads. For specific polymer transport properties, such as ion conduction, alternative coarse-
graining strategies have been developed that preserve a more direct connection to the underlying
atomistic picture. Maitra and Heuer[21] have developed a theoretical approach in which atomistic
simulations are used to simulate the short-timescale molecular processes associated with ion
conduction; these simulations are then used to parameterize a Rouse-type model for the long-
timescale polymer and ion conduction dynamics. Alternatively, recent extensions[22] of the
dynamical bond percolation model[23] employ short-timescale MD simulations to parameterize a
lattice-based model for ion transport in polymers, creating the possibility for efficient screening of
polymer electrolytes (Figure 2.6). Further development of coarse-grained models for the efficient
characterization and screening of candidate polymer materials is essential for discovery and
synthesis of new polymer materials.
Outstanding challenges for polymer
design. Diverse polymer properties of
interest and the combinatorial design space
for candidate materials: An important
obstacle in computer-aided polymer design
is that any particular materials application
rests upon multiple underlying polymer
properties, each of which presents unique
challenges for theory and simulation. To
aid the identification and screening of
polymer candidates, theoretical
frameworks are needed in which all
relevant polymer properties are adequately
modeled. Perhaps the greatest single
challenge for polymer design is the sheer
magnitude and diversity of candidate
Figure 2.6. For ion-transport in polymers, coarse-grained polymer chemistries, materials, and
models, such as the chemically specific dynamic bond conditions of interest. Any theoretical
percolation (CS-DBP) model, avoid long-timescale MD framework for predicting and screening
simulations by using short MD trajectories to parameterized candidate polymers must offer sufficient
efficient kinetic Monte Carlo (KMC) simulations.[22]
accuracy for the properties of interest,

37
along with the efficiency needed to cover sufficiently large swathes of the space of candidate
materials for a given application.
In summary, computer simulation strategies offer extraordinary practical benefits for the
design of new polymer materials by identifying promising candidate materials and by avoiding
wasted effort and resources on polymers materials unlikely to exhibit properties of interest.
However, computational polymer design and screening also offers profound theoretical challenges
associated with balancing the accuracy, efficiency, and generality of those theoretical methods. It
is expected advances in computing power, machine learning, and simulating coarse-grained
modeling strategies will lead to an even more central role for computation in discovery,
characterization, and design of next-generation polymer materials. The close collaboration of
theoretical and synthetic chemists will increasingly become a model for polymer research and
design.

2.2.4 To polymerize pre-organized monomers to create programmed architectures.


Motivation for developing new methods to polymerize pre-organized monomers. Nature has
exploited sequence control to store information in nucleic acids, structure and function in proteins,
and sensing in polysaccharides. Many of the techniques used by natural systems to achieve
sequence control also lead to well-defined and programmed molecular weights in which all chains
have the same length. Over the years, we have learned that most naturally-occurring biological
macromolecules are synthesized by a method that relies on pre-organization of monomers.
Pre-organization of monomers with external fields. With the available set of monomers and
polymerization methods, there is a largely unexplored opportunity to exploit external static and
dynamic organization fields to dictate primary chain structure. Examples of such fields include:
electric, magnetic, liquid crystalline, fluid flow, strain, or light. While many of these fields have
been explored to align block copolymers to control macroscopic morphology, there is a vast
potential here to control single-chain structure during polymerization.
These goals will require targeting specific monomer properties (such as magnetic
susceptibility, dipole moment, and polarizability) and mesogenic properties that respond to these
external fields to create a structured, crystal-like medium to dictate particular aspects of the
polymerization. Breakthroughs could vastly expand the monomer types beyond olefins to
acrylates, methacrylates, and styrenics, and use existing polymerization methods. Theoretical and
computational chemistry tools are needed to predict the properties of the monomers and their
magnitude. This will enable field-driven stereochemical control of polymerization.
Template-driven monomer ordering. A template can help pre-positioning monomers with
the right orientation and order. Outstanding challenges include: can we hijack biological
machinery to template and pre-organize monomers to create synthetic polymers? A related
challenge: can we learn from biological systems and create synthetic templates to pre-organize
monomers? As an example, using a DNA template to achieve polymerization methods that are
scalable and have the same level of sequence and structure control as in biology, but with a wider
set of monomer functionalities. Sawamoto, et al., reported an iterative process where a growing
chain migrates to control propagation with addition of a single-monomer at a time by combining
free radical polymerization and metal-mediated reversible cyclization chemistry.[24] To meet this
grand challenge, many of the following goals must be met: 1) design of monomers with encoded
functionalities for template recognition, 2) design of synthetic macromolecules as templates
containing information, functionality, and organization that directs subsequent polymer growth in
a programmed amplification of sequence, 3) use of both covalent and supramolecular bonding to
create complementary template-monomer binding that can dynamically/reversibly dictate the
38
primary structure, and 4) potentially solid planar or topological surfaces with nanoscale chemical
codes to spatially control/direct polymerizations.
Polymerizations in confined spaces. Spatial confinement of monomers and their orientation
during polymerization can lead to structural control typically not achievable via conventional
solution, melt, or gas phase. Can we derive from current concepts the effect of confinement on the
crystallization behavior of organic compounds? Can we translate them to monomers and to
structural control of polymer chains? It is likely confinement can minimize the chain-termination
pathways during chain-growth polymerization and narrow the dispersity. These confined
environments can be created with self-assembled block copolymer morphologies (such as
micelles, vesicles, and unilamellar lipid bilayer-like assemblies) or within the microchannels of
porous materials (such as silica, zeolites, molecular sieves, and 2D materials). Can the confinement
also provide encoded cues for atomic-level ordering of the monomers, for example, in a metal-
organic framework? Alternatively, can nanoporous materials be exploited to allow molecular
weight control by polymerization of ordered linear arrays of monomers in confined pours?

2.2.5 To program higher order structure formation (tertiary, quaternary) into


polymerization chemistry design for in situ access to complex nano-objects.
In nature, the behavior and function of biomolecules is inextricably linked to tertiary
structure (how individual chains fold into three-dimensional structures) and quaternary structure
(how multiple chains fold into three-dimensional structures). These higher levels of structure arise
at the monomer level as a consequence of primary (monomer sequence) and secondary structure
(hydrogen bonds between different monomer units influence structure). Efforts to control primary
structure in synthetic polymers should also develop toward elaborating on these methods to control
tertiary and quaternary structure.
Developing methods to synthesize polymers with controllable tertiary and quaternary
structure so the size and shape of the resulting polymer nano-objects can be controlled at the
nanometer level would have a significant effect in a range of applications (Figure 2.7). Such nano-
objects have the potential to influence protein-protein interactions, influence crystal nucleation,
act as synthetic antibodies, serve as
therapeutic drug delivery agents, or
function as new optoelectronic materials,
depending upon the composition and
functionalization of the constituent
polymers. If higher-order structure could
be programmed into individual polymer
chains so complex nano-objects could be
formed in situ or with minimal additional
processing, the effect would be
tremendous; it could allow rapid screening
Figure 2.7. Schematic illustration of designing higher order of multiple functional species for
polymer structure from a heterogeneous pool of monomers, optimization in specific applications.
from primary structure (sequence) through secondary
(intra-chain links) and tertiary (three-dimensional structure) Programming higher-order structure will
to quaternary (multi-chain three-dimensional structures). necessarily rely strongly on developments
in theory and in synthetic methods.
Tertiary structure in synthetic polymers is largely dictated by the rigidity of the backbone
used, but in specific classes of polymers (e.g., polyisocyanates, bulky methacrylates), secondary
and tertiary structure is determined by well-defined interactions between monomer units. In these
39
cases, the structure-determining functional groups are present at every repeat unit, so precise
control of primary structure is not critical to higher-order structure formation. At longer length
scales, crystallization is important in many polymers. In some cases, it forms relatively regular
higher order structures. There are opportunities for growth in this area with peptide-mimetic
polymers (peptoids, etc.) and in the resurgent interest in bottlebrush polymers and copolymers; but
expansion of the palette of structure-determining monomer units would be useful.
It is only in limited contexts that rational strategies for controlling polymer quaternary
structure are available. For example, block copolymer structures in bulk and solution can be
designed by understanding how non-covalent interactions drive intermacromolecular interactions.
Block and random copolymers with monomer units that can specifically interact in an
intermolecular manner have been used to encourage the formation of coarse quaternary structures
in polymer blends. Polymerization-induced self-assembly (PISA)[25] (in which non-covalently
assembled block copolymer nano-objects can be prepared in situ during polymerization) has been
used to synthesize a range of assemblies with quaternary structures determined by degree of
polymerization of a solvophobic block; these are typically limited to the canonical spherical,
cylindrical, and vesicular structures formed by amphiphiles (Figure 2.8A). The elaboration of
covalent polymerization with concurrent non-covalent polymerization provides another route
toward controlling higher-order structure (Figure 2.8B).[26] The growing body of work in the
design and synthesis of foldamers provides an invaluable starting point. Though much of this work
has been carried out on low molecular weight oligomers, the relationships developed between
abiotic monomer structure and larger scale structure will aid the design of larger structured
polymers (Figure 2.8C).[27] The tendency for chains of chiral monomers to form stereocomplexes
with chains of their enantiomers can also be exploited to form quaternary structures.
In biomolecules, tertiary and quaternary structures are stabilized by non-covalent linkages
(typically intra-/inter-chain hydrogen bonds) and covalent linkages (most importantly the disulfide
linkages in proteins). Covalent cross-linking has been used to stabilize tertiary and quaternary
structures in synthetic polymers (Figure 2.8D).[28] Selective cross-linking of structured bulk
materials, followed by dissolution, has been used to make structured nano-objects. For example,
selective cross-linking of the central polydiene block of polystyrene-block-polybutadiene-block-
poly(methyl methacrylate) in a phase-separated film, followed by dissolution, was used to prepare
non-symmetric Janus particles (Figure 2.8E).[29] Incorporation of complementary hydrogen-
bonding moieties along polymer backbones has been used to drive the formation of secondary and
tertiary structure in several polymer systems (Figure 2.8F).[30] The continued elaboration of these
stabilization strategies and the development of new strategies for preparing polymer nano-objects
will help prepare new generations of polymer-based materials, especially if robust, rapid, and
scalable methods can be developed.
Developing methods to construct relatively simple nano-objects is the likely first step toward
elaborating more complex design strategies. Common biochemical motifs—helices, sheets,
hairpin turns—have relatively well-understood design elements already explored in synthetic
polymers, though further honing of these systems will be necessary. In simple systems, it is
desirable to include modular elements that could further modify nano-objects through non-
covalent chemistry or click chemistry. Another obvious class of nano-objects for initial
investigations is that in which the size, structure, and symmetry is largely determined by the
structure of the parent polymer(s). For example, spherical objects can be prepared from tertiary

40
structures such as random coils in solution and quaternary structures such as spherical micelles
and vesicles.

Figure 2.8. Selected examples of strategies for the preparation of synthetic polymers with higher order (tertiary,
quaternary) structure. A. Polymerization-induced self-assembly (PISA) as a route to micelle-based quaternary
structure; B. Tandem covalent and non-covalent polymerization for the preparation of fibrillar quaternary
structure; C. Helical tertiary structure from a dendronized conjugated polymer; D. Intramolecular cross-linking as
a route to tertiary structure; E. Defined non-symmetric quaternary structure through cross-linking and dissolution
of phase-separated block copolymers; F. Introducing tertiary structure through selective intra-chain interactions
(compiled from [25-30]).

2.2.6 Accelerate the development of new polymers by discovering new mechanisms that
allow preparation of new polymers from readily available monomers.
There is a need for methods that allow production of novel polymers using new combinations
of readily-available monomers. Such approaches would reduce time to impact for new polymeric
materials.
The main types of polymerization processes and mechanisms studied and used commercially
today show specificity towards particular monomer families (olefins, dienes, styrenics, acrylates,
etc.). Mechanisms that can polymerize or copolymerize two or more monomers from different
families should be developed. This would enable a broad palette of new polymers exhibiting
enhanced properties compared to available materials. In contrast, it is more common to circumvent
mechanical limitations by using protecting groups (which require removal after polymerization
and add synthetic complexity and cost).
The most important commercial monomers or classes of monomers used in polymerization
processes include: simple acrylates and methacrylates, acrylic acid, acrylonitrile, butadiene,

41
isoprene, ethylene, propylene, vinyl chloride, vinyl acetate, simple higher alpha-olefins,
isobutylene, styrene, acrylamide, epoxides; and the building blocks for polyesters, polycarbonates,
polyamides, silicones, and other high-volume condensation polymers (Figure 2.9). All of these
monomers have been in commercial use for decades. Their synthesis, cost, purification, volume,
handling, and supply chain considerations have been optimized and are well understood. Thus, the
creation of novel materials from new combinations of familiar monomers is an attractive approach
to a new generation of polymers.
Polymer chemistry’s most important
classes of polymerization mechanisms all
exhibit limitations on application to
different families of monomers. Free-
radical polymerization offers little
mechanistic or microstructural control
during polymerization and is not applicable
to many high-interest vinylic monomers.
Similarly, anionic and cationic
Figure 2.9. A small group of monomer classes are used to polymerization mechanisms are limited to
make traditional high-volume commercial polymers. a subset of the most readily-used,
Developing new polymerization processes to copolymerize commonly-available industrial monomers.
new combinations of familiar monomers—whose handling,
supply, and cost are already optimized—would offer an
Yet these mechanisms are not useful for or
advantage to new polymeric materials. tolerant of many other readily-available,
commercially-relevant monomers.
Regarding coordination-insertion metal-mediated polymerization processes, mechanisms
have been developed which cover many but not all of the common monomer classes—olefins,
vinylaromatics, dienes, cyclic esters and oxiranes, acrylates, etc. Coordination-insertion processes
offer excellent potential for precise control of polymer architecture (tacticity, comonomer content,
regiochemistry, endgroup structure) through manipulation of spectator ligands and the use of chain
transfer agents. However, the majority of these mechanisms strongly prefer insertion of a particular
monomer type and do not tolerate crossovers between monomers with different structures.
Similarly, metathesis polymerization of olefins (e.g., Ring Opening Metathesis (ROMP) and
Acyclic Diene Metathesis Polymerization (ADMET))[31] has enabled the synthesis of many
unique and previously inaccessible polyolefinic, functionalized, and multiblock structures. But
metathesis polymerization processes that would allow for copolymerization with other potentially-
useful, industrially-relevant unsaturated monomers are not well-developed.
Advances and limitations in the coordination-insertion polymerization of olefins using
molecular catalysts illustrate the need for more generalized and tolerant copolymerization
mechanisms. The ability to copolymerize simple olefins with polar monomers in a high-efficiency,
low-cost process has been an unsolved challenge of catalytic polymerization science for many
years. The nonpolar, nonfunctional character of polyolefins presents fundamental property
limitations for this commercially-important class of polymers—for example, lack of adhesion and
compatibility with other polymer types. In the last 10 years, several important advances have been
made towards this goal using late transition metal catalysts.[32] There should be further
development of new catalysts with greater ability to copolymerize simple olefins with other
important monomer types while allowing for single-site-like activity and fine control of polymer
microstructure and architecture.

42
Completely new polymerization mechanisms that are not inherently biased or pre-optimized
for a particular monomer structure are needed. This alternative could enable previously-
inaccessible copolymerizations of common monomers. These mechanisms are likely to be
developed through translating and optimizing advances in catalytic organometallic (or organic)
chemistry into polymerization space. For example, the irreversible switching of a polymerization
process involving a single catalyst from preferring one type of monomer to another is common.
Previously unusable combinations of common monomers may become usable by developing
catalysts capable of reversible mechanistic switching during polymerization—polymerizing one
monomer class, then another, in sequence—to generate multiblock or other polymer structures.

2.2.7 To develop versatile methods for sequence control and determine how sequence
affects function.
Control of monomer sequence in synthetic polymers has advanced rapidly in the past decade
through ever-improving synthetic methods for greater control over primary structure.[33] For
example, living polymerization and alternating polymerization mechanisms have provided access
to block copolymers and periodic polymers, respectively. Ultimately, the sequence control of
nature’s biological machinery far surpasses what synthetic polymer chemists are capable of. While
there is inherent scientific challenge in synthesizing a polymer that approaches the complexity of
a biological polymer, there is also a need to understand why such a structure would have broader
impacts.
Key challenges to address are: (i) the design of new, inexpensive, and scalable syntheses of
sequenced polymers; (ii) the identification of fundamental structure/property relationships; (iii)
the development and incorporation of predictive theory; (iv) control over increasingly complex
sequences. Advances in these areas could move polymer synthesis toward addressing the grand
challenge of precision control over polymer structure and function.
The complexity of copolymers can range from statistical randomness to unique and precise
placement of thousands of monomers (Figure 2.10). In most cases, as sequence complexity
increases, the scalability decreases. At one extreme, the modification of biological systems with
recombinant genes or use of the ribosome allows synthetic polymers with the complexity of
biological polymers. However, when compared to statistical copolymers or commercial block
copolymers, the scalability of such biological methods is poorly positioned to have a societal effect
or to probe structure/property relationships. Solid phase synthesis with reagents eases the
difficulties and provides excellent control but consists of multi-step processes. This limits their

Figure 2.10. Schematic representation of macromolecular structures with varied levels of sequence
complexity.[34]

43
scalability and molecular weights. Advances in chain-growth and step-growth polymerizations
have enabled more scalable approaches to periodic polymer sequences of intermediate complexity.
Iterative exponential growth has been used with click reaction mechanisms to prepare periodic
polymers that are sequence- and stereo-defined polymers with periodicity.[35] Other periodic
polymers have been efficiently prepared via the pre-synthesis of ABC-type monomers that can be
enchained with either living or step-growth polymerizations. One of the more robust methods for
preparing sequenced polymers is use of polymerization mechanisms that exhibit kinetic
dispositions for alternating monomer enchainment. Exploiting relative reactivities allows for
perfectly alternating AB copolymers to be prepared in a single step and on practical scales. Another
strategy entails pulsing monomer addition with a favorable kinetic profile to give rise to polymer
chains containing discreet regions with comonomer incorporation at controlled distances of
separation.[36] Expanding these strategies to encompass increased levels of sequence complexity
is a critical challenge for synthetic polymer chemistry.
Even if more general and robust methods for synthesizing sequence-defined polymers are
discovered, there is a challenge to understand the importance of a sequenced polymeric material.
The structure/property relationships of random and block copolymers are increasingly well
understood in terms of their thermal and chemical properties. Extending these relationships to
sequenced polymers and exploring more advanced properties (information storage, ion transport,
conformation, etc.) is a challenge. Furthermore, even if such fantastic properties are realized, the
field must identify the practicality and reasons for pursuing such properties in light of advances in
competing materials (semiconductors, metal organic frameworks, block polymers). It has already
been shown that sequence can affect thermal properties such as glass transitions and melting
temperatures; chemical properties such as solubility and hydrophilicity; degradation rates of
biodegradable polymers such as poly(lactic-co-glycolic acid) (PLGA);[37] the conformations of a
polymer as a folded or self-assembled chain; and the ion transport properties of such synthetic ion
channels. These studies should be expanded to include other primary structures and more complex
sequences, but a deeper understanding of how and why these properties emerge as a function of
sequence is necessary.
Polymers with sequence complexities beyond block copolymer architectures form complex
structures which are a challenge to predict and design. Foldamers are oligomeric in nature and do
not behave like polymeric macromolecules but do demonstrate a sequence-dependent
supramolecular structure. Polypeptoids are suited for sequence-specific synthesis and subsequent
assembly into hierarchical structures. The precise placement of ionic groups throughout a polymer
chain via ADMET or ROMP has profound effects on the subsequent thermal and ion transport
properties of the polymer. The ability to model and predict these structure/property relationships
is a challenge requiring collaboration between theoretical and polymer chemists. Furthermore, the
challenge of linking structure/property relationships will require synthetic advances in preparing
increasing levels of complexity on scales permitting physical studies.
As the complexity of polymer sequence increases, the number of structural combinations,
and potential properties increases exponentially. Thus, it is a continuous challenge to obtain greater
precision and degrees of complexity via synthesis. The introduction a wide range of monomers
with precise placement without the need for step-wise synthesis is a challenge. It is also critical to
develop synthetic techniques that work with many polymer backbones and not exclusively with
the products of click reactions to compare properties of sequenced polymers to those that are
traditionally used. Additionally, it is a challenge to prepare sequenced polymers consisting of
monomers that are typically introduced through different polymerization mechanisms.

44
In summary, to advance sequenced macromolecules toward a societal effect, the field must
develop robust synthetic methods, understand the structure/property relationships of sequenced
polymers, develop predictive models, and synthesize increasingly complex primary structures.

2.3 Recommendations
1. Develop methods to synthesize macromolecules of any primary structure (sequence,
composition, tacticity, branching, architecture, topology, dimensionality [1, 2, 3-D],
functionality, MWD…) with reasonable time frame and cost.
2. Encourage the polymerization of renewable monomers under the principles of ‘green
chemistry’, to create cost-competitive, high-performance materials that are
environmentally benign; or that can be recycled by reversion to monomer or reformed
(vitrimer); or that simply last forever (durable).
3. Establish a theoretical framework for screening polymer targets that is (i) reliable for a
diversity of properties, (ii) atomic-scale in resolution to account for the effects of subtle
structural changes, and (iii) efficient enough to explore huge swathes of chemical space
and provide useful answers rapidly.
4. Polymerize pre-organized monomers to create programmed architectures.
5. Program higher order structure formation into polymerization chemistry design for in
situ access to complex nano-objects.
6. Speed development of new polymers by discovering new mechanisms or combinations
that allow preparation of new polymers from readily available monomers.
7. Develop versatile methods for sequence control and determine how sequence affects
function.

45
SECTION 3: HIERARCHICAL STRUCTURES AT MULTIPLE LENGTH
SCALES AND ENERGY SCALES
Discussion Leader
Rachel Segalman, UC Santa Barbara

Group Members
Stephen Cheng, University of Akron
Thomas Epps, III, University of Delaware
Jodie Lutkenhaus, Texas A&M University
Bert Meijer, Eindhoven University of Technology
Murugappan Muthukumar, University of Massachusetts Amherst
Bradley Olsen, Massachusetts Institute of Technology
Virgil Percec, University of Pennsylvania
Sam Stupp, Northwestern University
Karen Wooley, Texas A&M University

Special Acknowledgement
Gabriel Sanoja, University of California at Santa Barbara, graduate student
participant

3.1 Introduction
Polymers are unique materials capable of simultaneously achieving order over a wide range
of length scales. These hierarchical assemblies form due to an interplay between thermodynamic
and kinetic driving forces acting through the atomic (monomer), nanoscopic (sequence, chain
shape), and mesoscopic length scales. These forces ultimately determine the macroscopic
functional properties of a polymeric material. This assembly process, analogous to a chemical
reaction, has a complex free energy landscape with multiple local minima that can be accessed in
a controlled manner via processing pathways that depend strongly on external fields (e.g.,
temperature, stress, electric, magnetic). Importantly, order at every length scale affects both the
structure and the function of the resulting polymers.
The classic example of self-organizing hierarchical structures formed by polymers is how
even the simplest polymer, namely polyethylene, organizes into a crystalline state. Upon cooling,
entangled polyethylene chains in a molten state form macroscopic spherulites with a hierarchy of
structures (Figure 3.1). Many essentially two-dimensional thin lamellae (~10 nm thickness and
~10 micron width) bunch up into three-dimensional structures where the lamellae are twisted like
ribbons exhibiting gigantic chirality, although the polyethylene molecule itself is achiral. The
thickness of each lamella is orders of magnitude smaller than the molecular length, and each chain
traverses among several lamellae, resulting in a semicrystalline state. The semicrystalline
polymers, with the combined properties of mechanical strength and elasticity, have spawned the
modern Polymer Age. But a full understanding of how such hierarchical structures emerge remains
elusive, despite seven decades of effort.

46
Figure 3.1 Hierarchical assembly of flexible polymers into semicrystalline morphologies: (a) Polyethylene
crystallizes as spherulites, where multiple lamellae are bunched into three-dimensional assembly, with amorphous
regions intervening lamellae. Achiral polyethylene molecules exhibit gigantic macroscopic chirality in the form
of banded spherulites (b) seen in optical microscopy due to twisting of lamellae (c). Polyethylene crystallized
from solutions exhibits a plethora of morphologies: (d) flat hexagonal lamellae, (e) flat lozenge shaped lamellae,
(f) hollow pyramid, and (g) globules. Scrolls (h) and twisted lamellae (i) are also prevalent. Individual lamellae
are made of multiply folded chains, with crystalline bulk and amorphous interfaces.[1]

Even when intermingling and entanglements among chains are absent, as in crystallizations
from dilute solutions, polyethylene organizes into a plethora of morphologies. It forms flat
hexagonal lamellae, lozenge shaped lamellae, hollow pyramids, and globules, depending on the
crystallization conditions. Such morphologies are common for other polymers, too. Scrolls and
twisted lamellae are also common, depending on the nature of the polymer. Understanding the
kinetics of this organization process is rich and presents many challenges. For example,
semicrystalline polymers order better upon melting, and a molten polymer remembers previous
semicrystalline states during further crystallization. The fundamental conceptual framework
behind the spontaneous selection of hierarchical morphologies and their kinetic pathways lies in
the antagonistic interplay between the capacity of polymer chains to possess huge conformational
entropy—opposing order—and the short-range attraction among polymer repeat units—

47
facilitating order. The chain connectivity naturally results in topological defectsthat interfere with
ordering and promote hierarchical assembly.
Expansion into polymer architectures beyond homopolymers such as polyethylene has
brought tremendous opportunities to formulate many hierarchical structures with tunable
functional properties. As an example, block copolymers with interactions governed by short-
ranged dispersion forces form nanometer scale domains that align macroscopically to achieve
long-range order. The ability to control the properties of these building blocks to induce
hierarchical order and control polymer macroscopic properties has enabled applications such as
thermoplastic elastomers, semiconductors, and ion-conducting solid-state electrolytes. These
applications have largely relied on polymers in thermodynamic equilibrium or kinetically-trapped
nanostructured morphologies. Extraordinary challenges and opportunities arise from innovative
new polymer materials with complex structures that arise due to interactions that range across
multiple time, length, and energy scales; and whose assembly and macroscopic properties are
dynamic and change under external stimuli. Non-covalent interactions (e.g., electrostatic,
hydrophobic, hydrogen bonding) are critical for developing stimuli-responsive, adaptive, and
reversible materials because the interaction energies that dictate the hierarchical structures respond
to multiple orthogonal handles on faster timescales than their covalent counterparts. Thus, the
molecular engineering of hierarchical assembled polymers will rely on rational design of chain
conformation, mesoscopic structure, and macroscopic function in an integrated manner. Finally,
taking inspiration from nature, we must begin to think of assembly not as a process with fixed
thermodynamic states defined by equilibrium but rather as a pathway with useful intermediates
that yield a richer and more complex phase space.
Grand Challenge: To make the next leap in designing functional materials, we must
understand, predict, and use molecular building blocks, sequence,
conformation, and chirality to direct mesoscopic structure at desired
time and energy scales to control macroscopic polymer properties.
We envision that such fundamental insight will lead to new polymer applications including:
• Novel processing methods for semicrystalline polymers with tunable mechanical
properties
• Retaining the biological functionality that is the result of hierarchical structure in a
polymer designed system
• Directed self-assembly of multifunctional building blocks
• Adaptive materials
• Materials with designed and anisotropic properties (particularly transport and
mechanical properties)

3.2 Fundamental challenges


Hierarchical structures resulting from tradeoffs between short- and long-range non-covalent
interactions in polymers can be used to rationally design materials with unique and finely tuned
macroscopic properties. Further, the chain connectivity of polymers results in long-ranged
topological correlations in space and time. The interplay between polymer connectivity and
segmental interaction energy results in cascades of transfer of entropy into enthalpy in the free
energy landscape for the ordering processes of polymeric materials. Over the last decades, we have
learned to incorporate and control non-covalent interactions at multiple length scales, and this is
leading to an expanded range of applications for polymers in electronics, medicine, and other

48
technologically-relevant fields. Current polymer design is often modular and based on
complementary properties of individual components (e.g., block copolymers, polymer mixtures,
and polymer-nanoparticle composites); yet unique properties also can emerge from assemblies. In
its simplest rendition, current state-of-the-art hierarchically-assembled polymers have properties
that are linear combinations of the constituent components. We anticipate the next step will be to
develop rational design rules for hierarchically-assembled polymer materials that have different
properties than the individual chains. Engineering polymer materials with these emergent
strategies will become increasingly important in the coming decade. This behavior is analogous
to the hierarchical structures that result from primary amino acid monomer sequences and provide
biomacromolecules (e.g., proteins) with outstanding functions. Moreover, it is well-known biology
uses many other motifs for directing structure (e.g., α-helixes and β-sheets) resulting from defined
primary amino acid monomer sequences. By coordinating this ensemble of non-covalent
interactions, living systems produce materials with remarkable mechanical properties (e.g., bone,
skin, shells, silk); biorecognition and catalytic specificity; and efficient energy harvesting that do
not exist in the denatured, unstructured state.[2] We anticipate that as polymer synthesis
approaches the complexity and sophistication of natural polymers, the need to better control the
assembly of these systems to achieve the efficacy of nature-based design becomes significant.
The physical insights provided by understanding the role of non-covalent interactions on
guiding the hierarchical assembly of polymers can be exploited beyond biological applications.
For example, the optoelectronic properties of a crystalline conjugated polymer are different from
those of an isolated chain, as the bulk charge mobility and bandgap are the result of non-covalent
interactions and can be controlled through the hierarchical structure. Another example is that of
polymer-nanoparticle composites that exhibit mechanical and electronic properties which are
generally a linear combination of the individual components. However, there is a nascent
opportunity to combine building blocks such that unexpected physicochemical properties emerge
as a result of hierarchical assembly (Figure 3.2). In this case, the incorporation of a diblock
copolymer based on semiconducting and ion-conducting blocks into an inorganic battery electrode
results in a dramatic improvement in both mechanical and ion transport properties.[3]

3.2.1 Next generation molecular and polymer building blocks for hierarchical assembly
The rational molecular design of hierarchical-assembled polymers for advanced functional
materials requires engineering and synthesis of polymer building blocks to control the
supramolecular structures that efficiently transfer and amplify the microscopic functions to
emerging macroscopic properties. Polymer hierarchical assembly is a multi-stage process in which
each individual stage contains different chemical and physical reactions (covalent and non-
covalent synthesis) at various length, energy, and time scales. Thus, as we move towards targeting
functions and properties that are the result of a hierarchical assembly, inverse design principles are
required. Echoing the Materials Genome Initiative launched by the U.S. White House in 2011, the
most effective approach to address this challenge is practicing a tentative retro-functional analysis
scheme, which is intended to be a function-oriented, modular approach for molecular design by
considering two important factors: (i) function of the building blocks and (ii) their structures. This
approach can be meaningful only when the functions of the building blocks are relatively
independent of each other and are additive towards a collective macroscopic property. For
instance, over proper arrangements of the structural protein domains with different function, new
protein functions can emerge. Hence, the basic philosophy underlying this approach can be
described to be that (1) it should be function oriented; (2) it should be modular and efficient; and
(3) the building blocks are independent of each other yet have well-defined molecular functions,

49
precise chemical structures (formed by covalent bonding), and preferred secondary/packing
schemes (formed by secondary interactions) to assemble into hierarchical structures.
In the last decade, directed self-
assembly (DSA) of self-assembling
polymers for lithographic applications has
achieved tremendous success, leading to
not only perfection of simple patterns but
the creation of complex structures by
taking advantage of surface interactions in
polymer thin films. Spatial control of
patterns on the nanoscale in a scalable,
parallel, and robust fashion now seems
within reach. Similarly, as the
semiconductor industry begins to
incorporate three-dimensional
architectures, a need to direct assembly of
the polymer patterns in three dimensions is
becoming evident. Moving forward, this
avenue of research will not only include
polymers of more complex topologies but
also more building blocks with non-
covalent interactions that can make more
complex architectures. Recent success of
this modular approach is shown in Figure
3.3, where a wide variety of molecular
building blocks assemble first into
mesoscale architectures which then sit on a
remarkably complex lattice.[4] These and
even more complex architectures pave the
way towards multi-compartment, multi-
layered, and aperiodic structures for
Figure 3.2. V2O5 battery cathodes become flexible upon the applications beyond lithography, including
addition of small amounts of P3HT-b-PEO. This polymer the synthesis of new materials of complex
simultaneously enhances the electrochemical performance geometry within the polymer domains.
and the mechanical properties.[3] This approach could provide components
for applications ranging across targeted
drug delivery, photonics, catalysis, and energy conversion. In terms of building blocks, we
anticipate these will need to be designed to not only maximize etch contrast for lithography but
also deliver payloads such as dopants or seeds for a secondary synthesis stage within the assembly.
There are some intrinsic weaknesses in this inverse design approach. First, a rough prior
knowledge about the relationship between structure and function should be in place to guide the
initial design. However, this is usually not the case in new materials. Second, structure and function
are never fully decoupled in materials. In fact, emergent properties that result from mesoscopic
structure are ignored in this formalism. The critical step towards successful hierarchical assemblies
that are more than the sum of the parts is computer simulations and calculations to predict useful
chemical structures; topological design of the building blocks, and emergent properties that are a

50
result of assembly. Indeed, the tools necessary to predict emergent properties pose a major,
fundamental challenge for multi-scale modeling.
The inverse design approach is,
however, expected to be critical for
polymer science as our ability to synthesize
complex building blocks improves with the
development of chemistries that allow us to
tune a versatile and extensive molecular
parameter space. For example, as we gain
precise control over monomer sequence,
there is an opportunity to control
hierarchical structure and macroscopic
properties to an unprecedented degree and
elucidate the role of sequence on structure-
property relationships. The simplest
renditions of sequence control, gradient
and tapered structure in a copolymers, are
Figure 3.3. An example of a building block approach to
technologically interesting as interfacial
highly complex architectures. A wide variety of modular modifiers, reinforcement agents, damping
building blocks can be assembled into mesoscale structures materials, and thermoplastic elastomers
which can then be modularly built into complex three- because precise sequence control enables
dimensional assemblies. A15 lattices (Frank Kasper phases) tunability over chain shape (secondary
similar to those shown on the right of this image have been
previously observed for diatomic metal alloys but have
structure) and thermodynamic and kinetic
recently been observed to result from all of the mesoscale driving forces leading to long range order
structures.[4] (tertiary structure). Similarly, conducting
polymers of interest for organic electronics
exhibit optoelectronic properties which can be tuned with precise control over monomer sequence.
Given that charge delocalization and mobility are tied to chain shape and packing on the nanoscale
while interactions with light (absorption and emission) depend on mesoscale order, a deep
understanding of how to simultaneously direct primary, secondary, and tertiary structure is critical
to further development. Analogous challenges arise with increasing complexity in polymer
architectures and topology (e.g., linear, branched, brush, cyclic, star), which have the potential to
greatly broaden the array of hierarchically-assembled polymer materials.
Thus, a theory challenge is built into inverse design. Theory has made tremendous strides in
describing the equilibrium thermodynamics of assembling polymers but has not yet delved into
the kinetics of assembly (and the determination of kinetically controlled patterns); the molecular
design of the building blocks; and many of the non-idealities intrinsic to advanced functional
materials. For example, the process of polymer crystallization is extremely complex and
kinetically controlled (far from equilibrium). The classical view of polymer crystallization has
hinged upon a generalization of theories of crystallization of small molecules based mainly on
energetic considerations without accounting for chain connectivity. Aided by computer modeling
and sophisticated field-theoretic techniques, there have recently been attempts to address the role
of conformational entropy of chains in polymer crystallization. As an example, in the primordial
stage of polymer crystallization, each chain first forms baby nuclei with local smectic order, which
are connected by chain connectors (Figure 3.4a). As time progresses, the baby nuclei grow and
compete among themselves, mediated by the conformational entropy of chain connectors,

51
eventually forming lamellae. Polymer crystallization is thus a hierarchical process mediated by a
multitude of entropic barriers, as depicted in Figure 3.4b, where the computed free energy
landscape is plotted against the lamellar thickness and the local segmental order parameter. The
major effect of finite contribution of chain entropy is to stabilize the semicrystalline state in
equilibrium. The global free energy minimum corresponds to nanoscopic thicknesses of the
lamellae and not to the mesoscopic molecular length. As a result, the equilibrium melting
temperature does not correspond to the extended-chain state, as previously believed in the classical
literature. As the identity of the equilibrium melting temperature is crucial in dictating the driving
force for crystallization, the realization that the equilibrium melting temperature of lamellae is
controlled by chain entropy has opened a new direction of theory of polymer crystallization.
Computer simulations and theory also show the growth kinetics of semicrystalline polymers are
also controlled by entropic barriers arising from restricted chain conformations that are unique to
polymers and not present in small molecules (Figure 3.4c). Despite these recent advances, a
comprehensive theory and understanding of polymer crystallization under quiescent and flow
conditions do not exist.

Figure 3.4. Computer simulations reveal molecular mechanisms of polymer crystallization: (a) birth of baby
nuclei in the primordial stage[5], (b) free energy landscape of polymer crystallization showing the global stability
of finite lamellar thickness instead of extend chain dimension[6], (c) entropic barriers for the growth kinetics of
polymer crystals[7].

In principle, the initial selection of building blocks would be inspired by theory and iterated
with experiments until convergence to an optimum high-performance material with desired
macroscopic physicochemical properties. This iterative approach between experiment and theory
also will provide insight into issues such as the role of sequence and chirality of the formation of
hierarchical assemblies. Special attention should be directed towards integration of bonding on
multiple energy scales to create hierarchical structures (i.e., covalent and non-covalent bonding).

3.2.2 Thinking of assembly as a reaction: Non-equilibrium assembly and pathways


A particularly exciting possibility builds on the fact that, by directing the assembly of
polymers, one might be able to routinely access non-equilibrium states of matter, thereby
expanding considerably the palette of morphologies—and functionality—that is available for
applications. Recent proof-of-principle demonstrations with simple diblock copolymers suggest
that this is feasible: directed assembly under chemo- or grapho-epitaxy can be used to trap
metastable morphologies reproducibly; and experimental and theoretical evidence suggests the

52
same is true of controlled solvent annealing. To capitalize on the promise of non-equilibrium
assembly, not only for block copolymers but also for other forms of directed assembly, it is
essential that efforts be made to understand at a fundamental level the spatiotemporal evolution of
assembling polymeric materials over a wide range of scales. Several decades of research have
sought to predict the rheology of entangled polymeric materials, and it is only recently that
breakthroughs have started to emerge for simple homopolymers. Building on such successes,
theoretical and computational models must now be developed for not only the rheology of more
complex materials but for other path- and time-dependent properties (such as the transport of ions
within polymeric systems and the role of branching on semicrystalline morphology). Such models
will necessarily be expected to capture time-dependent morphological changes in the presence of
short- and long-range interactions; supramolecular interactions; and, importantly, in the presence
of applied fields (including stress, electric, electrochemical, magnetic, thermal) and temporal
exposure to other external cues.
From an experimental perspective, a shift in thinking must occur to conceptualize synthesis
and assembly in an iterative fashion. If approached in this manner, synthesis or chemical reaction
can be incorporated within a processing scheme in which bonds may be made or broken during
assembly. This also opens the door to co-assembling very heterogeneous components. As these
processing schema become more complicated, we must learn to harness the inevitably non-
equilibrium pathways to achieve desired structures. In this area, applications have outpaced our
understanding. For example, while many new applications, particularly in organic electronics,
require casting polymers from solvent or solvent annealing, solvent drying in even simple systems
is very poorly understood and leads to highly non-equilibrium structures. Experimentally,
differentiating long-lived and reproducible non-equilibrium structures from actual equilibrium is
not immediately necessary for applications but is critical to extending our designs to the next
generation. Experimentally, this will likely require fundamental, patient work to rationally turn
interactions on and off to understand the free energy landscape iteratively. Similarly, while
equilibrium structures are robustly described by existing simulation tools, new methods for
understanding non-equilibrium pathways are needed. As another example, polyelectrolyte
complexes and multilayers are inherently non-equilibrium structures. The nature of a complex
depends on simple processing pathways such as the order of mixing. Equilibrium-type complexes
can be achieved by prolonged annealing, either by the addition of salt or cyclic pathways.
However, even these materials can exhibit distinct properties subject to processing specificity. For
instance, the ionic strength and valency of the salt counterions strongly affect complex structure
in solution. The nature and dynamic evolution of non-equilibrium versus equilibrium
polyelectrolyte complex assemblies is not at all understood.
New theoretical paradigms are demanded by the omnipresent non-equilibrium states in
polymer self-assembly processes. These paradigms should go beyond the currently-practiced
theoretical methodologies, such as the self-consistent field theories and extensions of the Flory-
Huggins theory. The challenge resides on the calculation of the free energy landscapes for the
assembly processes and their temporal responses to externally- or self-generated stimuli. The
generic nature of an assembly process with many macromolecular components is equivalent to a
series of chemical reactions among small molecules, except that the topological correlations of the
macromolecules quickly push the macromolecular assembly into many long-lived, non-
equilibrium states. In addition to the quantum and classical forces operative at the sub-nano scales
associated with chemical reactions, conformational entropy associated with the very nature of

53
macromolecules amplifies the spontaneous emergence of non-equilibrium states in
macromolecular assembly.
Every step in the process of macromolecular assembly can be treated as a chemical reaction
with a collision frequency between two macromolecular reactants and a transition frequency
between the initial and final states. The capture of one macromolecule by another requires
knowledge of the pair potential between the corresponding molecules in the dielectric medium of
the assembly process. The calculation of the pair potential between macromolecules—ensuring a
robust representation of monomer interactions and polymer sequences, especially in polarizable
dielectric media—is presently a huge challenge. This challenge must be met in efforts to
understand the capture frequency of reacting macromolecules.
The transition from the macromolecular reactants to the final state is a dramatic
reorganization of polymer conformations, with even this one elementary step of self-assembly
involving multiple metastable states. The natural consequence of multiple ways of pairing among
monomers of the reactant molecules is the topological dereliction in the free energy landscape
associated with this step (e.g., Figure 3.4b). Unlike in the equilibrium considerations using
statistical mechanics, all intermediate metastable states are not equally accessible due to the self-
imposed pathways connecting them. Different pathways lead to different final states, and these
states are not inter-convertible. The choice of intermediate states that are thermodynamically stable
at every instance of the macromolecular rearrangement process does not necessarily lead to the
final state associated with a global free energy minimum. It is necessary to controllably unleash
external stimuli at appropriate times in the assembly process to make the final product. (These
external stimuli are not unlike the role of chaperones in evolved natural systems.) Such challenges
involving capture between two macromolecular reactants and the topological dereliction in their
structural reorganization in forming the final state get even more amplified when the product reacts
with another reactant, and so on. Systematic theoretical and computational approaches towards a
fundamental understanding of the connected pathways among metastable states in the free energy
landscapes of macromolecular assembly are critically needed to complement the experimental
advances being made in a broad spectrum of applications.

3.2.3 Active/Adaptive hierarchical structures


Current use of hierarchical structures generally follows a linear path from synthesis to
assembly to applications. That properties result from a hierarchy of length scales invites the
development of adaptable materials whose hierarchical assembly can be changed, on command or
during use, to switch properties (akin to biochemical cascades in living cells). We currently do
things that approach this but not on the timescale or with the triggers we need. For example, turning
ionic or electronic conduction off by dis-assembling a transport pathway would yield an adaptable
membrane material. Indeed, exactly this sort of adaptation is already possible using temperature
as a state variable on the very slow timescale of polymer diffusion (Figure 3.5). Materials that
adapt upon exposure to light or from mass diffusion of a chaperone molecule are alluring for these
applications.
New and advanced polymers should be designed with hierarchical structures and
macroscopic properties that can be modulated by external fields (e.g., electric, magnetic, thermal,
and stress). For example, electrochemical potentials are commonly harvested in biology to induce
changes in the hierarchical structure of the extracellular membrane of neurons and facilitate the
transport of neurotransmitters. Non-covalent interactions are essential to the design of responsive
hierarchical structures. For attractive interactions, a broad range of binding energies among
monomeric components is possible to control the time scales of response to external stimuli.

54
Figure 3.5. Reversible patterning of soft conductive materials by stimuli-responsive supramolecular
assembly/disassembly correlated with sol-gel phase transitions.[8]

Repulsive interactions of electrostatic origin have been shown to control hierarchical organization
among supramolecular polymers approaching macroscopic length scales; and in hybrid
organic/inorganic materials, these repulsive interactions have been shown to trigger significant
dimensional changes that are thermally reversible. One could envision examples in which rigid
covalent and supramolecular polymers are chemically designed to experience repulsive
interactions driven by multiple stimuli, including changes in pH, light exposure, and mechanical
forces. Further, layering a variety of orthogonal interactions (for example, hydrogen bonding,
ligand-metal interactions, molecular steric forces, and block polymer segregation) is essential to
creating several length scales of order and disorder that can be turned on and off individually.
Using orthogonal interactions to manipulate hierarchical structure and, therefore, active and
adaptive behavior will be most exciting if relatively small molecules in solution or in the gas phase
could be used to morph structures (Figure 3.6).
This approach will enable changes in materials or in properties that are not hampered by the
diffusion barriers. (Diffusion barriers would only enable changes in active and adaptive materials
to operate in extremely long time scales.) Furthermore, the multi-scale, spatial-nature
characteristic of non-covalent interactions allows bulk and solution hierarchical assembly in a fast,
reproducible, and reliable manner. For example, electrochemical potentials induce changes in the
hierarchical structure of biological membranes that allow ion-transport. Similarly, external fields
(e.g., thermal) induce changes in the hierarchical structure of proteins and let them adapt to retain
activity under conditions different from the native state.

3.2.4 Bio-inspired and biological assembly


Biology is often a source of inspiration, with its specifically-adapted functions emerging
from complex, hierarchically-structured materials. A grand challenge in polymeric materials
design is to replicate this level of complexity. Advances have closed in on this goal from two sides.
On one side, synthetic polymer systems have grown increasingly more complex, including
controlling primary monomer sequence and polymer architecture and structuring these polymers
into hierarchically-ordered materials. From the other side, as the biotechnology revolution matures
and our mastery of biology grows, biology itself has become a tool of the polymer scientist. As
shown in Figure 3.7, assemblies of amphiphiles mimic biological membranes and enable precise
delivery of drugs, proteins, and other agents in a biological environment. Making highly precise
vesicles of predetermined size and shape is one of the successes of bio-inspired patterning
approaches.[10, 11]

55
Figure 3.6. A macroscopic example of a hierarchically adaptive material is a polymer gel with responsive pore
size coated onto a fabric, which is then integrated into a cloth/garment. The gel responds by collapsing (right
image) to selectively close pores (cross-link) in a fabric in response to only toxic organophosphate agents. This
illustrates a high selectivity, adaptive response to change both structure and properties.[9]

Continued advances in molecular biology, industrial biotechnology, and bioconjugate


chemistry make it easier and more affordable to produce precisely-engineered biopolymers (i.e.
DNA sequences, designed oligopeptides and peptoids, and artificially-engineered protein
polymers) at a scale suitable for materials engineering, as shown in Figure 3.8. In this example,
fully folded proteins are directed into assembly in much the same way that traditional block

Figure 3.7. Polymersomes and dendrimersomes are examples of how synthetic chemistry and hierarchical
structuring are yielding assemblies of complexity approaching that of biological systems. The use of biohybrid
polymers functionalized with peptides or polysaccharides can introduce biological functionality into the
assemblies.[10, 11]

56
Figure 3.8. A) Self-assembly of fully-folded globular proteins into nanostructured materials that show primary,
secondary, and tertiary structure preserved within mesostructured materials. B) Sequence engineering of elastin-
like artificial protein polymers controls their phase separation pathway, leading to kinetically-arrested gel
structures with high modulus and toughness. Left: [12]; right: [13].

copolymers can be controlled.[12] Indeed, detailed sequence engineering of the proteins can even
provide additional handles with which to control assembly.[13] These polymers provide
unprecedented opportunities to use sequence-specific primary sequence to attain precise control
over hierarchical structure formation. They have inherent advantages due to their biological
functionality, which makes them particularly suitable for biomedical materials, biocatalysis, and
sustainable polymer applications. Biological systems also provide a special opportunity to
understand the role of molecular dispersity, as many biopolymers are naturally perfectly
monodisperse, providing ideal models. Beyond materials engineering, studies of natural materials
and structures directly relate to NSF Grand Challenges in understanding the mechanistic origins
of disease or the origins of life, which often requires a detailed knowledge of the hierarchical
structures present in living systems.
Recent successes in de novo design of proteins provide inspiration. The ultimate goal is to
design a monomer sequence and develop processes that translate this control over primary
sequence into control of secondary, tertiary, and quaternary structure. This feat is much closer to
achievable for biopolymers (protein and DNA) than for synthetic materials or even
polysaccharides; it requires substantial collaboration between theory and experiment for its
realization. For example, this understanding directly informs the design of artificial and active
enzymes and of synthetic vaccines made from hierarchically-assembled materials. Similarly,
hybrid biotic/abiotic systems required detailed design-based understanding. These hierarchical
assemblies provide a direct route to nanopatterning proteins and enzymes at high densities while
also preserving high levels of protein activity. By effectively nanopatterning proteins, these
systems could aid the fabrication of highly-efficient biocatalytic, bioelectronics, and biosensor
devices. Efforts should be directed to developing thermodynamic and kinetic models for this kind
of complex hierarchical assembly.
The fundamental principles of building hierarchical structures are always the same whether
a system is synthetic or biological; therefore, we can elucidate these common principles and
develop a unified approach to designing hierarchical structures. Progress will be fastest when study
of synthetic and natural polymers is united; biopolymers have much to contribute to our
understanding of classical problems in polymer physics such as gelation, reptation, lower critical
solution behavior, and hierarchical assembly. We have some idea how to design primary and
secondary structure in biopolymer systems but less in synthetic systems. In all systems, de novo

57
design of tertiary and quaternary structures is in its infancy. For example, our understanding of
chirality and how it affects the structural hierarchy is poorly developed. As shown in Figure 3.9,
chirality appears to transfer upwards in length-scale from a chiral monomer to a chiral
nanostructure,[12] but homochirality has also been observed from completely achiral building
blocks.[13] The polymer community is uniquely situated to provide the models that will inform
biological discovery, not just take inspiration from biology.

Figure 3.9. Chirality has transferred across lengthscales in chiral bio-inspired systems (top) but also originates
from completely achiral building blocks (bottom). The former case raises intriguing questions about chirality
transferred across lengthscales, while the latter suggests opportunities to address NSF Grand Challenges, including
the breaking of symmetry that appears to be at the mechanistic origin of life (Top: [14]; bottom: [15]

Due to the complexity of biopolymers and biohybrids, advancing our understanding of these
materials demands particularly close collaboration between theory, simulation, and experiment.
Understanding electrostatics in crowded environments, the effect of water, and ion solvation are
all emerging challenges for natural and synthetic systems. Advances in coarse-grained modeling
will have a central role, but many effects that are critically important in biopolymers (such as
polarizability and chirality) will have to be treated at the atomistic or quantum level. This will
require linking theory and simulation across time and length scales, an additional ongoing
challenge that may require experts from outside the polymer community.

58
3.3 Key enabling technology needs
As more complex materials are developed to achieve more highly-controlled and complex
structures, new synthetic, computational, and characterization tools will be needed. They should
rapidly and iteratively explore the relationship between molecular structure, processing, final real-
space structure, and properties. Increasing material and structural complexity simultaneously
places increasing demands on instruments. They must measure more structural features and do so
more quickly, generating tremendous challenges. Further, in situ measurement of structure-
property evolution during a given response to stimuli is still needed. New materials feature
complexity on many different length scales arising from the use of new monomers with new
specific interactions, monomer sequence control, non-idealities in chain conformation, new chain
architectures, and interactions. Further, new materials may exhibit multiple solvent, temperature,
and other field and chemical-responsive interactions; tuning the relative dependence of each
component of the molecular structure to these interactions can dramatically affect the structural
and processing landscape of the materials. Importantly, characterization tools used for feedback of
molecular design and processing must yield reproducible results of the structures in the final, post-
processed state; measured structures must not be inconsistently biased from the true structures by
processing conditions such as staining and fluorescent labeling, which require the incorporation of
additional components into the material.
Furthermore, as hierarchical materials are developed for specified applications, the relevant
structure will frequently exist only under specific environmental conditions (for example: in
aqueous settings, humidity effects; in the presence of specific biological components, pH
variations; in operando devices such as batteries, etc.). Therefore, it is critical that the environments
relevant for critical applications are anticipated that characterization tools are developed that are
capable of completely probing relevant structures in those environments, and that those tools are
available for researchers. Simultaneously, it is also critical that basic tools increase in accessibility.
Tools such as neutron scattering, X-ray techniques, and advanced TEM (see also Figure 3.10) are
readily accessible, primarily at national laboratories; it is critical that collaboration and
accessibility to such workhorse instruments be further increased. Ultimately, increasing
complexity in molecular structure and final structure will require increasingly automated methods
in synthesis, processing, and structure-property characterization. Such automation will achieve the
control and rapid, iterative feedback critical to actually realizing the potential of these materials.

3.3.1 New tools needed for hierarchical structure and properties (synthetic,
characterization, modelling, and computation)
To achieve materials with the desired hierarchical structure and properties, new synthetic,
characterization, modeling, and computational tools will be required. New synthetic tools will be
critical to achieve the degree of both incorporated functionality and iterative control needed to
rapidly explore molecular structure space. To understand the effect of precise changes in molecular
structure, the material must be reproducibly synthesized while changing only the desired property.
New, robust synthetic methods for highly controlled materials have progressively been introduced;
in particular, these have recently included sequence-controlled materials via solid-state platforms
(especially for biological and biologically-inspired polymers) and solution-state, iterative
protection-deprotection schemes. However, improved and controllable techniques must
incorporate new functional classes of materials into complex molecular structures. For example,
conjugated and electrically active polymers cannot be synthesized at the level of control of
biological materials or materials amenable to anionic or reversible-deactivation radical

59
Figure 3.10. Characterization of block copolymer solution assemblies. Schematic showing (a) a subset of the
structural information obtainable from various scattering techniques and (b) a subset of images extracted using
different microscopy techniques for a spherical polymer micelle. Panel (a) highlights common structural
dimensions that can be readily probed in an idealized spherical micelle, along with associated reciprocal-space
scattering techniques; these dimensions include Rc (core radius), Rg (radius of gyration), Rh (hydrodynamic
radius), R (micelle radius). Panel (b) illustrates complementary information that can be gleaned from real-space
microscopy imaging. Many of these techniques can be incorporated into in situ characterization setups.[16]

polymerization techniques. Development of new synthetic methods must also clearly address
purification schemes to achieve the required level of purity and degree of dispersity for repeatable
behavior; small amounts of remaining monomer, homopolymer, small molecules, or salt may have
large effects on the resulting structure and properties.
Finally, there is tremendous opportunity for theory to expand from predicting equilibrium
states for complex materials to predicting the path-dependent, non-equilibrium structural evolution
of materials (i.e., the processing-dependence). Ultimately, theory should determine critical
measurable properties of hierarchical systems that form excellent estimates of the structural states
of systems during processing; systems can then be developed to control processing and navigate
the complex energy landscape towards the desired final state. This goal moves beyond predicting
the equilibrium structures that arise from complex polymers; it will require advancements from
theory of non-equilibrium structures in polymer systems. Experimentalists, theorists, and
instrumentation experts must collaborate and develop the required measurements and control
systems.

3.3.2 Testing of properties over relevant length- and time-scales and orientations
To readily understand hierarchical structures, examine the structure and properties of
hierarchical structures over the range of relevant length and time scales. Local structure often
results in features translated to larger scales (for example, chirality transfer from the monomer or
small molecule level to nano- or micro-scale structures, or the combination of properties achieved
by nanostructuring in thermoplastic elastomers). Electrical transport is similarly highly dependent
upon molecular and structural features over a wide range of length-scales, ranging from the chain
structure and local chain packing to grain structure and degree of material that participates in
percolated high-mobility pathways. Furthermore, hierarchical and functional materials often are
fundamentally anisotropic, with resultant anisotropic structure and properties; testing methods
must be able to examine these materials in the relevant geometries and orientation. It also will be
critical to understand the interaction between material structure, local material dynamics, and the

60
net time-scales of materials response. For example, responsive materials or biological sensors must
have a net response significantly faster than the time-scale of change in the property they are
measuring, controlling, or responding to; materials design must understand how the time-scales of
the desired property relate to the local dynamics of the designed hierarchical material.
Rheology, dielectric spectroscopy, and NMR- and neutron-scattering-based dynamics
techniques can capture materials dynamics over a range of time-scales and interpret them as motion
over a range of length-scales. Precise understanding of even simple materials via these techniques
remains limited. It is further complicated by the limited sample quantities available for iteratively-
synthesized, highly-specialized hierarchical materials. For instance, rheo-SANS scattering is a
powerful technique that can reveal the otherwise-inaccessible interplay between rheological and
structural properties under flow in real time, but due to the necessary sample quantity per
experiment (several milligrams to grams), dynamic studies of polymeric materials like protein
bioconjugates or responsive polypeptides are often precluded. We believe the development of new
techniques capable of probing structures and properties over an expanded range of length- and
time-scales will be critical to fully understanding next-generation hierarchical materials, especially
for limited sample quantities. This polymer characterization goal is central to effectively designing
molecular building blocks that direct mesoscopic structure and assembly for macroscopic
properties on relevant time- and energy-scales.

3.3.3 For hierarchical assembly experiments: can we analyze structure without stains or
tags and within in-situ environments?
New hierarchical assemblies will likely have highly-specific structures critical to their
properties. Their structures will often be designed to be stimuli-responsive, demonstrating
significant conformational change from environmental perturbations (temperature or pH variance,
introduction of trace small-molecule additives, etc.). Therefore, rigorously analyzing structures
becomes increasingly important. Heavy-metal stains or fluorescent tags for conventional
visualization techniques may interfere with imaging of actual structures. Therefore, imaging
techniques that can differentiate regions of separate chemical identity without artificial tagging
will become invaluable for increasing our understanding of hierarchical assemblies. Current
techniques at chemical resolution grant significant promise (for example, EDS-based TEMS
techniques, imaging via techniques such as X-ray microscopy and STXM); however, these
techniques require specialized materials preparation, are highly destructive, or lack resolution.
Specifically, many newly-developed X-ray techniques feature extremely high photon fluxes,
which yields tremendous capabilities but also challenges soft materials research.
New synchrotron and other high-flux techniques need to be developed with specific
instrumentation and user protocols for soft-materials work. Ideally, these procedures will help
users find the conditions needed to optimize meaningful data collection with minimal materials
destruction, thereby increasing user productivity with valuable beam time. In addition, as
processing or examining new materials increasingly require unique/transient environments to
demonstrate their final properties , it becomes critical to study materials structure in those
environments. Techniques such as environmental SEM and cryo-TEM are capable of imaging
chemical specificity of surfaces or structure in hydrated environments, respectively. However,
these laborious techniques lack resolution and only provide a snapshot of the sampled assembly in
time; tools must be developed to rapidly and iteratively probe these structures. More generally,
environmental stations that can track in-situ structural evolution in controlled environments must
continue to be developed and maintained. Some materials must be used in inherently
multicomponent environments in which they interact with a population of other materials.

61
Studying changes in the target materials and local structure uniquely from the surrounding
components will require continued development of experimental technology and of more
sophisticated modeling and theory. Observation of the structure-property evolution in operando,
such as for a solid polymer electrolyte during battery operation, presents further challenges in
instrumentation, as these experiments should ideally “see through” the target device.
Finally, it is important to develop tools which can specifically be used to understand the
changing structure of materials during the process of assembly. These materials will use the
environmental chambers described above to probe the path-dependence of material structural
development on changing conditions and rates. Structures often contain multiple critical features
that cannot be probed by a single technique; by incorporating multiple complementary techniques
into a single measurement setup, the overall state of the structure can be probed during the
structural evolution. An example of this is QCM-D with simultaneous FTIR, in which changes in
polymer swelling, modulus, and chemical signature are simultaneously monitored. Such
instruments will either need to be designed to be modular (to easily switch out instruments for
different materials systems), or they will need to be designed by scientists who will predict which
complimentary sets of instruments will be most crucial for studying the next generation of
hierarchical materials.

3.3.4 Core and regional facilities


Highly specialized characterization equipment is often available only at world-class facilities
(for example, the national laboratories). While these facilities are critically important for enabling
previously inaccessible scientific advances, the existence of limited numbers (often merely 1-2 for
a particular instrumentation) in the United States or even the world means extremely limited time
can be allocated to any given scientist. Without extremely high-throughput systems, these
instruments are often not designed to support the rapid feedback and iteration required for
understanding and developing structural control over hierarchical assembling materials. Therefore,
it is important to develop new programs, that enable workhorse instruments to achieve rapid
structural feedback. This should be done by developing regional centers, more widespread
adoption of lab-scale and benchtop instruments, and excellent high-throughput, mail-in programs.
Standardization of access programs across facilities will also improve progress; shared user
agreements that give users access to key capabilities in complementary facilities would
dramatically improve the speed at which scientists can understand the results.
Reciprocal space techniques remain a core need for characterization in polymer science.
They are complementary to real-space imaging techniques, which can rarely access the smallest
structural length-scales in polymeric materials. Techniques such a small- and wide-angle X-ray
scattering and neutron scattering are central to reciprocal space structural characterization; NMR
is critical to probing molecular structure. These apply not only to structural but also to
characterization of dynamics; neutron and NMR-based techniques remain some of the more
important ways to probe polymer dynamics. Further, neutron scattering makes it possible to design
specific contrast into materials via deuteration; similarly, the presence of elements may be
leveraged for enhanced X-ray contrast near absorption edges via techniques such as RSoxS.
Maintaining scientists’ access to these techniques and continually improving their resolution and
sensitivity will be important for advances.
Tools for modeling and data analysis must be developed and continuously improved. Next
generation data analysis tools should contain tutorials for data verification from particular
instruments (i.e., ensuring statistics are appropriate, backgrounds are subtracted correctly, and the
user is notified of anomalies). Software should also robustly perform reduction and modeling. As

62
scientists increase the throughput of samples, it will be important to automate more of the data
reduction and analysis while also improving the expertise among users of the techniques,
emphasizing the strengths and limitations of models. Perhaps most importantly, standardized and
accepted confidence intervals for fits of data to models should be defined; given that for some
techniques, many curves can ‘look’ the same, such standardization of confidence intervals will
provide guidance and uniformity in the field. Finally, instruments must support high-throughput
testing, and modeling and data analysis must be increased to process the results quickly and
reliably.

3.3.5 Robust theory and modeling are integral to characterization—required to do more


detailed data analysis (dynamics from NMR, hierarchical structure analysis,
understanding of transport)
Theory and modeling are critical to (1) understanding what structural behaviors are possible,
(2) knowing which experiment is ideal for demonstrating unique behaviors relative to models, and
(3) interpreting that experiment. Thus, theory and modeling can provide vision, planning, and
interpretation for the experimental community. For theoretical work to inform experiment,
significant understanding, close collaborations, and dialogues must exist between each
community. They must understand the capabilities and challenges faced by the other. Further, the
interface between theory and experiment represents an important educational challenge. Many
experiments are not complicated enough to require a specialized PhD student to develop a model
for them, yet they are distinctive enough for ‘one-size-fits-all’ software co be incapable of
interpreting the results. It is important to give students the tools to fill this niche themselves.
More advanced models may involve highly optimized methods in the code base and are often
maintained by individual groups. While re-developing all of this for a new group or student may
be impractical, other scientists may want to use or these models or expand on them. If theory
groups can share their well-documented, well-maintained methods and let other researchers
modularly extend and add capabilities while referencing the original work, the accessibility and
usefulness of these models will be dramatically increased. Again, as these groups invite others to
work with and build on their models, they must explicitly describe key skills needed before
beginning; they must develop tutorials for components specific to the model. To share components
of models, ideas, and methods between groups (both between theory groups and with
computational groups), it will become increasingly helpful for computational methods to be built
on a standard modular platform so software components can be built, modified, and added without
restructuring the rest of the work. Modularity, where possible, will let theory development become
increasingly faster, more adaptable, more accessible, and more collaborative. This becomes more
critical as experimentalists approach more diverse chemistries, architectures, processing methods,
and environments.
Scientists will use these modeling skills and tools to predict, parameterize, and describe the
expected behavior and the experimental response of a system for a given chemistry, architecture,
and inputs. It also will let scientists understand what aspects of the system model correspond to a
measurable property: designing for a given property is not useful unless scientists can confirm
experimentally that the property is displayed. Models should be more regularly used to predict not
just whether the quantity will be measurable but also whether the signal will be “loud enough” to
pick up and under what conditions such signals will be optimized; this grants the experimentalist
insight into the experimentally tunable behavior. When more than one state may yield the same
measurement, models help yield insight into what optimal combination of additional
measurements is necessary to identify a unique behavior. This framework can be used to inform

63
the development of both experiments and experimental techniques to determine what experiment
or combination of experiments will be most meaningful and efficient.

3.4 Recommendations
Overarching challenge: to make the next leap in designing functional materials and systems,
we must understand and use monomer structure and sequence, molecular conformation and
chirality, and mesoscopic and macroscopic structure. Thesew should be used simultaneously on
desired time- and energy-scales to control properties. Extraordinary challenges and opportunities
occur when desiging novel polymer materials with complex structures that arise due to interactions
ranging across time-, length-, and energy-scales; whose assembly and macroscopic properties are
dynamic under external stimuli.
1. Direct the assembly of inherently complex and multi-functional building blocks. This
includes retaining the biological functionality that results from hierarchical structure in
a designed system. We can make a macroscopic heart valve that retains its macroscopic
functionality but not the mesoscopic functionality of a cell or the microscopic
functionality of enzyme. Similarly, the transfer of chirality and other attributes from the
molecular to the macroscopic length-scale presents a largely untapped opportunity to
control structure and properties.
2. Directed self-assembly techniques have made great strides in lithographic applications,
but further advancements require directed assembly in three dimensions.
3. Uses of hierarchically engineered materials treat the structure as static. There are great
opportunities in patterning dynamic architectures to make materials that can change
properties on demand or sense and heal their own defects. This will enable new
paradigms in design of materials for switches, membranes, electronics, biological
materials, and structural components.
4. Non-covalent interactions (e.g., electrostatic, hydrophobic, hydrogen bonding) are
critical for developing stimuli-responsive, adaptive, and reversible materials because the
interaction energies that dictate the hierarchical structures respond to multiple orthogonal
handles on faster time-scales than their covalent counterparts.
5. We must begin to think of assembly not as a process with fixed thermodynamic states
defined by equilibrium but as a pathway with useful intermediates that yield a richer and
more complex phase space. Assembly can be designed similar to chemical reaction paths,
with many unit operations taking advantage of the kinetics and the thermodynamics of
assembly.

64
SECTION 4: INTEGRATED MEASUREMENT, ANALYSIS, AND
PREDICTION
Discussion Leader
Chinedum Osuji, Yale University

Group Members
Coray Colina, University of Florida
Russell Composto, University of Pennsylvania
Jane Lipson, Dartmouth College
Timothy Lodge, University of Minnesota
Christopher Ober, Cornell University
Rodney Priestley, Princeton University
Richard Register,Princeton University
Thomas Russell, University of Massachusetts
Christopher Soles, National Institute of Standards and Technology

Special Acknowledgement
Yekaterina Rokhlenko, Yale University, Student Participant

4.1 Introduction
Advances at the frontiers of polymer science are indivisibly linked to the ability to measure
and predict structure, dynamics, and properties with appropriate fidelity. In turn, the utility of data
generated by computational efforts and by traditional laboratory experiments relies on the
availability of appropriate methodologies for analysis. Computational efforts and laboratory
experiments are producing increasingly large and complex datasets. This brings associated
challenges regarding information classification and storage, and timely data processing. A tight
integration of measurement, analysis, and prediction is needed to catalyze advances in the next
decade.
Grand Challenge: To develop new tools—experimental and computational methods and
theory—and leverage them in an integrated way across any relevant
length scales and time scales to accurately measure, predict, and
elucidate underlying connections among structure, properties, and
dynamics of any macromolecular system.
The physico-chemical behavior of polymers involves a broad range of time and length scales.
Further, they are often out of equilibrium. Their behavior is, therefore, a function of their history
and is subject to evolution over time. Indeed, the structure and properties presented by
macromolecular materials and systems are functions of their thermal and mechanical history. As
a result, processing plays a substantial role in determining their structure and properties.
The pathways along which structure and properties evolve during processing are complex.
Elucidating these pathways and unraveling the complexity to generate a robust understanding of
structure-property-processing relationships is a key goal for the field. The ability to measure,

65
analyze, and predict structure and properties during processing is therefore of considerable
importance.
The impressive advances made in recent decades in experimental methods for characterizing
structure and dynamics should not mask the significant challenges that remain, particularly at small
length and time scales and during processing operations. A variety of goals can be articulated in
this regard. For example, knowing the position of “all the atoms at all times” would be a substantial
leap in our ability to describe the structure, dynamics, and properties of macromolecular systems;
even a static description of the precise chemical structure or composition of polymers at the single-
chain level would be tremendously enabling. Precise characterization of monomer sequence
represents a compelling goal in polymer science, with important connections to analytical
chemistry. Further, characterization of structure, dynamics, and properties across multiple length
and time scales remain key concerns. Concurrently, researchers must capitalize on the
sophisticated tools for characterizing polymer structure, dynamics, and properties, and must
continue developing such tools. Critical goals for the next decade include: ensuring access to
facilities and proper user training; and the availability of infrastructure to support manipulation,
storage, and analysis of large and complex datasets.
Modern science relies on theory and simulation to provide important insights about the
structure and properties of materials. Theory and simulation may aid in rationalizing and analyzing
complex experimental datasets. We also look to theory and simulation to provide robust
predictions of structure, dynamics, and properties. Further, we seek quantitative agreement
between experimentally-measured quantities and their predicted values. However, it is not yet
possible to correctly predict the behavior of polymers in bulk, in confinement, and in mixtures
with other polymers from first principles calculations or from simulations. Progress in this area
would greatly improve our ability to design macromolecular materials in a rational way to provide
desired structures, dynamics, and properties. Further, it would guide the design of processing
strategies such that the optimal intrinsic properties associated with a given polymer composition
could be realized.

4.2 Challenges, Needs, Opportunities


4.2.1 High resolution spatial and temporal mapping of structure, dynamics, and properties
The behavior of polymers originates in structures with characteristic length scales that span
several orders of magnitude. At one extreme, polymers are defined on sub-nanometer (nm) length
scales by the identities of their constituent atoms and their bonding topology. Statistical segment
sizes and persistence lengths are one level higher, with typical dimensions of one nm for flexible
linear polymers. The size of the polymer chain, as captured by the radius of gyration, depends on
the molecular weight of the polymer and can range from ~5 nm to ~50 nm for large molar mass
chains. Then, in semi-crystalline polymers and in block copolymers which undergo self-assembly
by microphase separation, there is structure on the length scale of tens of nm, due to the crystalline
lamellar thickness in the former case and the self-assembled nanostructures in the latter. Finally,
there is micron-scale structure in the form of chain-folded lamellae in semi-crystalline spherulites,
in grains of ordered block copolymers, and in component domains in phase-separated polymer
blends.

4.2.1.1 Precise characterization of monomer level structure


The properties of polymers depend on their structure and the nature of the interactions
between their constituents. In many cases, however, the details of the connections between

66
structure and properties are obscured by the inability to control and verify that structure at the
required length scales. This is true at the monomer level. There is a grand challenge centered on
the synthesis of polymers with precisely defined monomer sequences. The sophistication of protein
function in natural systems based on precise arrangements of amino acid sequences, as encoded
by DNA, provides inspiration in this regard. The sensitivity of block copolymer assembly to end-
group chemistry highlights the potential that monomer sequence and identity have to affect
polymer properties. Some encouraging successes on precise polymer synthesis have been
achieved. The coming decade will see intensified efforts to develop versatile synthetic strategies
(e.g., monomers, polymerization catalysts, and coupling schemes) to address the grand challenge
of enabling precise polymer synthesis. Advances in polymer synthesis have outpaced advances in
characterizing monomer sequence at the single-chain level. An associated challenge, therefore, is
the accurate determination of monomer sequence and topology in polymers. Indeed, the challenges
go hand-in-hand—accurate determination of sequence will be a critically important analytical tool
for efforts to control sequence precisely.
Characterizing chemical composition at the monomer length scale represents a considerable
challenge. Presently available methods, such as chemical spectroscopy (e.g., NMR) and
chromatography (e.g., size exclusion chromatography), provide statistical information about the
average local composition or size of ensembles of polymer chains. However, these methods are
ill-suited to the precise determination of monomer identity, stereochemistry, and sequence at the
single-chain level. New approaches are required. Advances in single-chain manipulation and
spectroscopic interrogation using, for example, scanning probes or microfluidic instruments may
become important in this regard. Determining monomer sequence and topology will likely require
synergistic efforts from polymer scientists, analytical chemists, and instrument developers.
The tools we use today to characterize, separate, and purify polymers cannot address the
needs of some emerging applications. In these cases, the functional properties of the polymer itself
may provide a more sensitive reflection of chemical purity and structural order than the analytical
tools can. For example, in semiconducting polymers, it is widely known that device mobilities or
device efficiencies can be improved by repeated purifications, often beyond the threshold of
analytical tools to detect the improvements in purity or related changes in structural order. New,
more sensitive analytical methods are needed.

4.2.2.2 Dynamics
The dynamics of polymers in solution and in the melt is a topic of fundamental concern.
There has been considerable success developing models that correctly account for the linear
viscoelastic properties of polymers in the melt and in solution. These have been extended recently
to the highly non-linear shear and extensional flow profiles characteristic of almost all commercial
processing operations.[1] This success is critically important to the manufacture and use of all
polymers, especially the so-called “engineering polymers” found in everyday products. The
dynamics of polymers in confinement and at interfaces, and the transport of charge carriers and
molecular solutes in polymers, are issues of contemporary and emerging concern. Successfully
addressing the associated challenges in measurement, analysis, and prediction will underpin the
development and use of new classes of functional polymer materials.
Charge transport in polymers is a topic of broad importance. For example, several battery
and fuel cell technologies rely on polymers as electrolytes. Charge migration in electronic
polymers is central to the function of organic electronic devices. Understanding how chemical
composition, processing, and structural defects affect the dynamics of ion migration and the
mobility of electronic charge carriers in these materials is important. It is also important to

67
understand the way charge carriers interact in these systems—how the ions interact with electrons
and the related changes in electrochemistry. Polymers that undergo controlled redox reactions
could play a significant role in advanced applications such as neuromorphic computing, and in
electrochemical energy storage.[2] In these instances, ion mobility alone cannot fully describe the
functional behavior of the polymer. Additional parameters (such as reaction rate constants and
chemically-specific dynamic measurements that can track electrochemical changes) are needed
and would provide transformative advances in these scenarios. High fidelity measurement of the
dynamics of ions in charge-containing polymers with sensitivity to their ionic state is a key goal.
Correlation of dynamics with structure and morphology is another. Connecting continuum-scale
measurements of conductivity to intrinsic microscopic properties such as diffusivity would be an
important outcome of such efforts.
A second area of concern is nanocomposites. Nanoparticles in a polymer can alter the
relaxation behavior of the polymer chains due to chain interactions with the nanoparticle surfaces,
even at relatively low nanoparticle loading levels. The way the chain dynamics are affected by the
nanomaterials is a complex function of several factors, including the type, size, and amount of the
nanomaterial present, and the degree of nanomaterial dispersion. Prediction of relaxation spectra
in inhomogeneous or otherwise nanostructured systems is an outstanding challenge.
Quasielastic neutron scattering (QENS) methods (neutron spin echo, backscatter, time-of-
flight) are important tools for characterizing dynamics in polymers from picosecond (ps) to
nanosecond (ns) time scales. Isotopic labeling allows selective examination of dynamics of
individual species in polymer blends or otherwise heterogeneous systems, such as nanocomposites.
However, QENS techniques lack sensitivity to the chemical state of the mobile species—for
example, in probing hydrogen dynamics, QENS cannot distinguish between the motions of free
water molecules vs. hydroxyl ions.
Chemically specific measurements of polymer dynamics are usually performed using NMR
methods (e.g., pulse-field gradient NMR, 2D-NMR to correlate dynamics of chemically distinct
species) and are expected to grow in importance. There have been advances in the integration of
NMR probes into scanning probe instruments that could be effectively utilized to provide spatially-
resolved data on polymer dynamics. However, NMR techniques generally lose sensitivity below
microsecond time scales.
There have been exciting developments in the field of 2D-IR (infra-red) measurements that
identify dynamic correlations between species with chemically distinct IR signatures on ns to ps
time scales. 2D-IR measurements have recently been used to study ion-containing polymers[3]
and water dynamics in gyroid surfactant mesophases.[4] They are expected to increase in
importance in the coming decade, particularly with improvements in the bandwidth (wavelength
range) over which dynamic correlations can be made.
X-ray photon correlation spectroscopy (XPCS) provides access to dynamics on time scales
from microseconds to several tens or hundreds of seconds and on length scales from microns to
nanometers. Resonant soft X-ray scattering (RSoXS) provides enhanced signal-to-noise in
scattering experiments and chemical specificity in structural data by tuning the X-ray energy to an
absorption edge of a species present in the system. The combination of XPCS and RSoXS has been
used to study dynamics in hard condensed matter. XPCS/RSoXS represents an opportunity to
elucidate dynamics in polymers with chemical specificity but without the need for isotopic
labeling. Resonant inelastic X-ray scattering (RIXS) combines the chemical specificity of RSoXS
with access to the same dynamical time and length scales of QENS. RIXS measurements therefore
have enormous potential for polymeric systems, especially in ionic and electronic polymers with

68
high-mobility charge states. The highly coherent, bright, soft X-rays needed to perform the
measurements above are becoming available at 3rd and 4th generation synchrotrons as our national
user facilities are upgraded and as new measurement capabilities come on line (e.g., NSLS-II).
Engagement between the polymer community and national user facilities will ensure the
availability and relevance of these measurement techniques for advancing the frontier of polymer
science.
Recently-developed, comparatively simple microscopy methods for studying dynamics also
represent opportunities for polymer science. Differential dynamic microscopy (DDM) is one such
method. DDM relies on computation of differences between images recorded at different times to
extract the intermediate scattering function. It provides information qualitatively similar to that
obtained in a dynamic light scattering experiment. A second method is interferometric scattering
microscopy (iSCAT), which uses a coherent light source to enable label-free detection and tracking
of nm-scale objects, such as nanoparticles in a polymer matrix (Figure 4.1). Direct particle tracking
at sufficiently high frame rates and appropriate spatial correlation of observed particle motions can
offer new insight into polymer dynamics, including questions about dynamical heterogeneity.

Figure 4.1: The need for speed: a) Tracking of CdSe rods (5 nm x 55 nm) in a polyacrylamide gel at fast (red)
and slow (blue) imaging rates. b) Histogram of log(D) at fast (red) and slow (blue) rates. Slower rates
underestimate the diffusion because they miss details of the rod displacement. Unpublished results by Composto,
Caporizzo and Goldman, University of Pennsylvania.

4.2.2.3 Structure characterization and chemical mapping


The increasing architectural and compositional complexity of polymers calls for improved
measurement and analysis methods for high resolution structure characterization and chemical
mapping. Atomic force microscopy (AFM) continues to be an important tool for characterization
of polymer microstructure, including imaging modes based on mechanical contrast, electronic
work function, and electrical conductivity. An emerging method, photo-induced force microscopy
(PiFM), allows nanoscale spectroscopic chemical imaging (Figure 4.2). The method uses
modulation of tip-sample interactions based on light-induced sample polarizability changes. As
with conductive AFM, PiFM can be used as a contrast mechanism, or it can be used to generate
an infra-red spectrum from regions as small as ~10 nm.

69
Figure 4.2. Infra-red spectra (hyPIR: hyperspectral PiFM infra-red) acquired using photo-induced force
microscopy (PiFM) of a sample consisting of nanoscale polystyrene droplets dewetted from a
poly(methylmethacrylate) film on a silicon substrate. The colored masks define the averaged spectra of the same
color in the plot. Image credit: Molecular Vista, Inc.

For ordered polymer structures, scattering by soft X-rays provides an important chemical
contrast mechanism; judicious selection of X-ray energies at the absorption edge of constituent
species in the polymer (Figure 4.3) is necessary. Soft X-rays from coherent sources also provide
unique opportunities for coherent diffractive imaging (CDI) methods, such as ptychography. CDI
is a lens-less technique that involves solution of the phase problem by combining iterative
algorithms and oversampling in various forms (viz., plane wave CDI; scanning or ptychographic
CDI; Bragg CDI). The resulting 3D reconstructions can have resolutions as small as ~10 nm and
feature the chemical specificity provided by X-ray absorption. CDI is an opportunity for polymer
science, if radiation damage is suitably mitigated. 3D imaging at nm-scale resolution with chemical
specificity would be highly relevant for a broad range of problems, for example, when studying
nanostructured or chemically heterogeneous systems. Appropriate tools and algorithms must be in
place to conduct the computationally-intensive analysis needed to reconstruct the image of the
system. An intriguing possibility is that
CDI may enable atomic-scale structure
determination of single molecules using
short femtosecond X-ray pulses to acquire
data before the sample is damaged by the
beam, i.e., using so-called “diffract then
destroy” schemes. Such pulses are
becoming available at 4th generation
Figure 4.3. Arrangement of microphase separated synchrotrons via free-electron laser
cylindrical microdomains of poly(1,4-isoprene) and poly(2-
vinylpyridine) in a polystyrene matrix of a poly(1,4-
sources.
isoprene)-block-polystyrene-block-poly(2-vinylpyridine) Direct visualization of dynamic
as deduced by resonant soft X-ray scattering. Three photon processes in polymers can provide
energies across the carbon 1s absorption edge were selected significant insight on unresolved problems.
to isolate the scattering contributions from each polymer. The direct microscopic visualization of
Experiments were conducted at beamline 7.3.3, Advanced
Light Source.[5]
self-diffusion by reptation in entangled
solutions of fluorescently labelled DNA

70
highlights this point. In situ transmission electron microscopy (TEM) offers the possibility of
direct visualization of processes at nm-length scales and time scales of ~1 ms and longer. It enables
structural changes to be observed under ambient conditions or while samples are subjected to
heating/cooling, mechanical deformation, photo-illumination, electric fields, etc. A recent study
demonstrates the utility of in situ TEM for soft materials by direct observation of diffusion of
polymer micelles in water.[6] Time-resolved TEM or dynamic TEM (DTEM) involves the
manipulation of a large pulse of photoemitted electrons to form high resolution images, offering
access to nm resolution and ns-µs time scales. DTEM is an emerging method and instruments are
not widely available. While DTEM has been principally applied to study transient processes in
hard materials, it can be adapted to biological samples and soft materials, and “diffract-then-
destroy” strategies, adopted from soft X-ray CDI, may help mitigate beam sensitivity issues. In
situ TEM and DTEM both represent unique capabilities which have not been extensively exploited
in polymer science. There are intriguing opportunities in the potential of these techniques to
address a variety of topics, such as structural evolution pathways in nanostructured polymers and
structure development during phase transitions.

4.2.2.4 Super-resolution optical microscopy and spectroscopy


Super-resolution microscopes are far-field optical microscopes, typically visible light
fluorescence microscopes with resolutions on the scale of tens of nanometers, i.e., beyond the
traditional diffraction limit of optical microscopes. With the circumvention of the diffraction limit,
there is a strong case to be made for applying the benefits of optical microscopy to polymer systems
with mesoscale structural order. These benefits include the ability to simultaneously and non-
destructively provide local 3D structural, multiphase (multicolor), dynamic, chemical, and
potentially, spectroscopic information. Opportunities and challenges related to these aspects are
briefly discussed below within the context of two popular implementations of super-resolution
microscopy—stimulated emission depletion (STED) microscopy; and single molecule localization
microscopy methods (SMLM) such as photo-activated localization microscopy (PALM), and
stochastic optical reconstruction microscopy (STORM).
The potential to image down to the macromolecular scale has been demonstrated, including
3D imaging of block copolymer thin films[7] and in situ images of living crystallization-driven
block copolymer nanostructures (Figure 4.4).[8] There is considerable potential to reveal details
of dynamic processes in polymers, as demonstrated in studies of nanoscale dynamic
heterogeneities in polymer solutions.[9] Imaging using highly photo-stable nitrogen-vacancy
centers in diamond has provided a resolution of 4 nm using STED microscopy; prototypical
commercial dyes offer resolutions ~ 20-30 nm. An important area of research therefore is the
development of high-performance fluorophores, as achievable resolutions are invariably limited
by fluorophore photo-stability and by brightness in the cases of STED and SMLM microscopy,
respectively. For polymer research, dyes must be developed that are photo-stable and bright and
that minimally perturb the structure and dynamics of systems under study.
An exciting prospect is the exploitation of fluorophores as site specific, i.e., local,
environmental probes in polymers. pH and voltage sensitive dyes are commercially available, and
photoacid distributions in polymers have been mapped. The orientation of individual dipole
moments could be used to gauge the orientation of nanoscale objects; dyes could report on local
variations in mobility or free volume. Additionally, individual molecules and therefore structural
features can be counted, although this is currently limited to tens of molecules in a diffraction-
limited volume. The integration of other spectroscopic label-free modes (such as coherent anti-

71
Stokes Raman spectroscopy and infra-red
spectroscopy) into super-resolution
formats represents a compelling prospect
for spectroscopic imaging of polymers.
Given the potential impact of super-
resolution microscopes, the question of
access is worth examining. Simpler, single-
lens systems are increasingly widely
available. By contrast, the advanced
double-lens, adaptive optics systems are
the purview of only a small number of labs
and are therefore likely candidates for
inclusion in mid-scale user facilities.

4.2.2.5 Unravelling complexity in


structure evolution
Figure 4.4. Top: TEM, confocal microscopy, and
stimulated emission depletion microscopy (STED) images
As noted earlier, polymers are often
of a poly(styrene-block-2-vinylpyridine) block out of equilibrium, and their properties can
copolymer.[7] The nanostructure of the system is resolved be strongly influenced by the manner in
in STED using a fluorescent dye which has been attached to which they are processed. The interplay
the poly(2-vinylpyridine) block. Bottom: Single molecule among structure, processing, and
localization microscopy (SMLM) mapping crystallization
driven assembly in solutions of a poly
properties is complex, even in polymers
(ferrocenyldimethylsilane- block-dimethylsiloxane) block and processing scenarios which might be
copolymer.[8] The polydimethylsiloxane block has been thought of as being relatively simple, and
conjugated with a fluorescent dye. TEM and wide-field that are widely used industrially. Such
fluorescent images are shown for comparison. The profile complexity is on display, for example, in
shows the improvement in spatial resolution of SMLM
compared to conventional wide-field imaging.
the interplay between structure and
mechanical properties in drawn semi-
crystalline polymers and in the
development of randomly interconnected microporous morphologies in polymer membranes
during fabrication by solvent-induced phase inversion. The level of complexity is considerably
higher in multicomponent systems that may also involve polymer crystallization and microphase
separation and which may feature polymers with novel chain architectures, for example, in the
processing of nanocomposites of semiconducting polymers. Unraveling the complexity of
structure and property evolution during processing is a challenge at the frontier of polymer science.
Addressing this challenge experimentally requires the development of quantitative multi-modal
metrology which can be applied in situ during processing operations to provide real time
information regarding local evolution of structure and properties. Integration of data from separate
(or combined) measurements using variable angle spectroscopic ellipsometry; grazing incidence
X-ray diffraction; and UV-Vis or infrared spectroscopy offers promise in this regard.
The challenge is also manifested in computational efforts where the need to bridge multiple
length scales and accurately represent multiple physical phenomena in simulations of polymers
during processing is computationally intensive. The development of powerful multi-scale methods
and efficient algorithms to provide more than phenomenological descriptions of structure-
property-processing relationships is needed. Efforts in this area are at still at an early stage. It is
expected that the need for advanced computational and metrology tools will grow as the

72
complexity of structure-property-processing relationships in functional polymers continues to
increase.

4.2.3 Advancing Structure, Property, and Event Prediction

4.2.3.1 Force field development and validation


Simulation methods have become increasingly sophisticated, able to handle larger systems
and describe longer length scales and time scales. However, to fully exploit the most recent
technical advances, there needs to be a greater emphasis on systematically confronting simulation
results with experimental data and much more effort on meticulous design of force fields that are
relevant to real systems and actual applications. These efforts should be advanced in a
complementary manner.
Consistent and explicit accounts in reported research of the details of force fields used would
help elevate the standard of results produced by computational efforts. Such details could include
the origin; initial intended application; past and current validation of force fields; the reasons for
force field selection; and the sensitivity of reported results to force field selection. Moreover, the
development of new and more accurate force fields for simulating the molecular dynamics of
polymers is a necessary and significantly enabling goal for advancing the frontiers of polymer
science and should be seen as such. Efforts along these lines are clearly warranted as there are
numerous scenarios in which existing force fields are inadequate. For example, there is high
interest in charged polymers confined by metal or high dielectric surfaces, yet simulation advances
are being hampered by lack of adequate force fields for such situations. The need for force field
development is also highlighted in the simulation of polymers in mixtures with ionic liquids.[10]
The demand for development of new polymers underlines the need for force field development,
i.e., to ensure theory and simulation keep pace with experimental progress.
Accurately capturing molecular geometrical characteristics and details of all relevant non-
bonded interactions for macromolecular systems is a non-trivial challenge. At many levels, force
field development calls for the availability of high quality thermophysical property data for
validation purposes. Increasingly, however, high quality data regarding fundamental
thermophysical properties of interest (e.g., density, thermal expansivity, virial coefficients) are
relatively scarce. Left unaddressed, the outcome will be less robust integration of theory and
simulation with experiment. This is particularly pernicious because current challenges require
exactly the opposite, and because technical developments on the computational side are opening
up new possibilities for accelerating the design of new materials by tightly coupling theory and
simulation with experiment. Greater support for collaboration between experimental and
theory/computational groups is needed. The benefits of such collaboration are visible in recent
progress using theoretically-informed, coarse-grained simulations to elucidate the three-
dimensional structure of block copolymers assembled on patterned templates (Figure 4.5).

4.2.3.2 Developing and using analytical theory


On the analytical theory side, there are some different challenges. It is often the case that
there are theoretical approaches available that provide insight going well beyond the most widely
used treatments favored by experimentalists. If theory, simulation, and experiment are to work in
tandem so as to move the polymer materials field forward, there needs to be recognition that
treating experimental data using phenomenological parameter-fitting methods does not represent
an advance in understanding of fundamental principles underlying behavior. Theorists must make
their methods more accessible and straightforward to apply and correspondingly, experimentalists

73
Figure 4.5. Three-dimensional structure of poly(2-vinylpyridine-b-styrene-b-2-vinylpyridine) films determined
from theoretically informed coarse grained simulations (left, middle) and from TEM tomography (right) for
different widths of the substrate guide stripes as indicated, in units of the lamellar spacing L0. Adapted from [11].

must be held accountable to using the most current theoretical methods for interpreting data.
Initiatives which lower the barrier for experimentalists to independently exploit state of the art
theory and simulation tools contribute positively in this regard, as recently demonstrated.[12]
Applications of microscopic theories invariably involve fixing the values of physically-
motivated parameters via experimental property data. Yet it can range between difficult and
impossible to find basic characteristic data on recently-developed polymers. For example,
densities, thermal expansion coefficients, and isothermal compressibilities are often essential
metrics needed to apply fundamental theory in a predictive way for bulk systems. This difficulty
echoes the issue addressed above in the context of force field development. In addition, these are
properties that can change for a material depending on its formulation (thin film, loaded with
particles, bulk, etc.); fundamental data covering a range of formulations is needed, but is almost
nonexistent. Efforts which better motivate experimentalists to undertake the required
measurements would help address this challenge, as would greater support for collaboration
between experimental and theory groups.

4.2.3.3 Thermodynamic interactions


The segment-segment interaction parameter, also known as the (Flory) χ parameter, is
ubiquitous in describing the thermodynamics of polymers in solution and in the melt. It is defined
strictly in a mean field context for components which display regular solution behavior; as such,
it is purely enthalpic in nature and a simple linear function of inverse temperature only. In practice,
it is used well beyond such bounds; it becomes, in effect, an excess free energy function. Therefore,
the effective parameter deduced from experimental data is always observed to display a substantial
entropic component.[13] For example, even blends of isotopically substituted polymers show a
significant excess entropy component to χ. Furthermore, χ is almost always composition-
dependent; it then becomes important to take into consideration which derivative of the free energy
is being probed experimentally. This situation is not unique to polymers—accurate descriptions of
phase equilibria in mixtures of even simple organic liquids rely on the use of local-composition
models for component activity that are more sophisticated than a single-valued interaction
parameter.
The persistence of overly simplified representations of thermodynamic interactions, while
convenient in some respects, is seriously limiting in others and should be addressed. The current
situation impedes quantitative use of thermodynamic data to design polymers for practical
applications, such as block copolymers for advanced lithography. Further, there is often poor

74
agreement in literature-reported values for interaction parameters deduced from experiments on
similar systems. The situation is further complicated by the lack of methods to directly calculate
interaction parameters for new polymer systems. The current state of the art must be advanced to
facilitate rational design of materials based on accurate representations of thermodynamic
interactions, both calculated and measured. The shortcomings of the current parameter-based
scheme could potentially be overcome by the use of a thermodynamic interaction function, which
reflects the multidimensional character of the thermodynamic interactions in polymer systems.
Advances will require concerted activities mobilizing computational and experimental
contributions and leveraging centralized resources for the reporting of thermophysical properties.
Quantitative agreement between carefully-conducted simulated and experimental neutron
scattering measurements of the Flory interaction parameter is possible, as demonstrated recently
by configurational bias Monte Carlo simulations using united atom force fields (Figure 4.6).[14]
The significance of this work is that detailed simulations using transferable potentials for
intermolecular force fields of oligomers may reveal the importance of different contributions to
the free energy of mixing. This represents a proof of concept that simulations based on chemical
structure alone may point the way towards an empirical “χ function” that could be used by
experimentalists to categorize phase behavior of polymer mixtures and block copolymers. From
the perspective of theory, many of these complexities could be avoided by using expressions for
the free energy of mixing that are (a) physically transparent, (b) sufficiently accurate for practical
purposes, and (c) dependent on only a few, independently accessible parameters. Examples of
these expressions already exist.

4.2.3.4 Managing property data


Advances in simulation and theory both demand the availability of high quality
thermophysical property data. Such data,
however, is often not available. A concrete
recommendation to address this gap is that
scientists who create new materials should
be strongly encouraged to provide
characterization data reflective of
thermodynamic and other properties. The
results should be reported in the literature
and also stored in a database or repository
accessible to the scientific community. The
repository should be administered by a
suitable national body, and the database
must be actively curated to maximize the
utility of information stored therein. The
results should be extensively annotated
with the metadata required to make it
useful to researchers. For example, a single
value for a polymer thermal expansion
Figure 4.6. The picture shows a simulation of atactic coefficient is not useful unless other related
polypropylene (yellow) and head-to-head polypropylene information is recorded. Examples of such
(red) at 440 K, 50% by composition, each oligomer has 12 information include the polymer molecular
carbons. Image credit: Q. Chen, T. P. Lodge, and J. I. weight (and, preferably, its dispersity or
Siepmann full molecular weight distribution function

75
and regiochemistry), the temperature range over which data were taken, and the pressure used
during the measurement. Standards will need to be defined by the community to ensure a uniform
set of common practices and understanding of terminology (e.g., how a glass transition
temperature should be defined).
The availability of plentiful data will provide opportunities for fundamental advances. For
example, if glass transition temperature values are recorded using four or five different
experimental methods for the same (or similar) samples of a material, it provides an opportunity
to test whether different techniques are probing the same physical response. If the same response
is being probed, it may be the case that the way that technique is effectively averaging over the
sample and the time scale of the experiment differs from method to method.[15] An abundance of
available data would motivate theory and simulation efforts to clarify the implications of accessing
different length scales or time scales in characterizing numerous properties and phenomena, e.g.,
the transition to glassiness and numerous other properties.
A repository for property data would provide a needed outlet for the accumulation and,
importantly, the organization of data from high-throughput characterization experiments. The
repository would enable researchers to use such data effectively to advance polymer science. An
effective repository for property data is one that is actively managed and readily accessible to
polymer researchers. Such a resource will significantly accelerate the discovery and development
of new polymer materials, an outcome that resonates broadly with the goal of the Materials
Genome Initiative.

4.2.3.5 Predicting maximal polymer properties


The intrinsic properties of polymers are dictated by their chemical composition and their
structure. As a practical matter, polymer materials seldom display their maximal intrinsic behavior
as their properties are degraded or otherwise altered from what would be expected based on
chemical composition and structure. The alteration of properties can come about due to the
presence of impurities or structural defects or simply lack of morphological control. For example,
when a polymer is tested for its photovoltaic performance, the measured “efficiency” convolutes
the behavior of many components of the overall system, and it is difficult to resolve how effective
a given polymer structure might be if all of the other attributes were ideal. Whether maximal
properties can be achieved in many polymers remains an open question, as processing
methodologies must be able to provide the required structure to achieve those properties.
Moreover, the maximal properties themselves are often unknown as they cannot be predicted with
great accuracy. Progress in this area has been limited. The available group contribution methods
and phenomenological models are not quantitatively accurate and often fail in the prediction of
properties of new polymers.
Advances in theory and simulation are required to provide useful quantitative predictions of
maximal polymer properties using a minimum number of assumptions. Developments of this
nature would be greatly enabling and would set polymer science and engineering on a trajectory
that has been traversed in other classes of materials. Examples of properties of interest include
mechanical metrics such as tensile modulus, toughness, strain-to-break, and yield strength;
transport properties such as ion, electron, hole, and dopant diffusivity; optical constants such as
refractive index, transparency, absorptivity, birefringence, and dichroism; and thermal diffusivity.
One could then envision theory and simulation-aided design of macromolecular materials with
novel, unprecedented functions—for example, low density, high stiffness, and high thermal
conductivity polymer fibers or polymer membranes in which the selectivity towards different gases
or solute species is predicted a priori with good fidelity.

76
4.3.1 Rare events in polymers
The behavior of polymers during routine use can be anticipated based on straightforward
characterization of their “average” properties. The point at which failure occurs is determined by
their ultimate properties but cannot be anticipated precisely. Although ultimate properties such as
yield stress or yield strain can be measured, failure in polymers is, in fact, a probabilistic event,
and we lack a detailed understanding of the manner in which it occurs. One may pose a simple
question—“Why does a polymer fail at a particular location at a particular time”. Post facto, one
can argue the failure occurred due to the introduction of a defect or to localized accelerated aging,
which decreased the durability of the specimen. However, we have limited to no ability to predict
the occurrence of the rare events associated with polymer failure.
Consider dielectric breakdown in polymers. A polymer thin film, an insulator, is placed
between two electrodes, forming a capacitor. The potential difference between the electrodes, or
average electric field in the polymer, is increased. Charge builds up on the electrodes as the
potential difference increases, until, at some point, an unknown event occurs, leading to an electric
discharge through the film; and failure of the capacitor as a permanent conducting pathway is
generated in the polymer. The pathway for this discharge can be spectacular (Figure 4.7), but we
do not know the nature of the initiating event nor can we say why the breakdown occurred at a
specific electric field strength. While statistically there may be an average breakdown field
strength, it is certainly not a deterministic predictor of failure from one film to the next. Yet, it is
precisely this breakdown that limits the ultimate lifetime of the material. Whether the process is

Figure 4.7. Visualization of electrical trees implicated in dielectric breakdown in low-density polyethylene. 3D
imaging was performed by X-ray ptychographic tomography with a resolution of 92 nm. Insets show
photomicrographs of the electrical trees. The area identified by the red circle in the top right inset is shown in the
ptychographic image. The red areas represent a 3D rendering of the iso-surface of the electrical trees formed
during breakdown. The electric field was applied in the direction indicated. Dashed box I shows 4 “prestep”
structures which are partially connected to the main tree by low density regions. The appearance of these prestep
structures suggests that the electrical trees grow in a step-by-step fashion. Adapted from [16].

77
electromechanical, where the electric field applies a strain on the polymer, or whether it is purely
electrical or mechanical is not understood. New light is being shed on this problem by recent work
which employed X-ray ptychographic tomography to provide 3D images of electrical trees in low
density polyethylene (Figure 4.7).[16]
Similar statements regarding the occurrence of rare events can be made with respect to
mechanical failure in polymers. While the shape of a polymer specimen can produce stress
concentrations, failure occurs on a much smaller length scale than the specimen itself, ultimately
involving the rupture of individual chemical bonds, Figure 4.8. One can ask why one specific bond
(and not another) ruptures, leading to the cascade of events that ultimately ends in macroscopic
failure of the polymer. There are a multitude of analogous situations in which rare events dictate
the ultimate properties or other behaviors of polymers. A quantitative understanding of rare events
that can account for the location and time of event occurrence would enable the development of
strategies by which polymer lifetimes could be improved or more accurately known.

Figure 4.8. Mapping of bond breaking during crack propagation in an elastomer using an in-situ optical imaging
technique. Incorporation of sacrificial chemo-lumninescent cross-linking molecules provides direct evidence of
the role of stress localized at the crack tip.[17]

It is anticipated that mapping structure, dynamics, and properties at suitably high spatial and
temporal resolutions will yield rich datasets that will significantly inform our understanding of rare
events. This anticipation rests on the hypothesis that the time and location of rare events is encoded
in the polymer by its microscopic history and that structural and dynamic heterogeneities play an
important role. Measurements will need to be conducted under relevant conditions, for example,
in polymers under mechanical strain or in the presence of strong electric fields. Observing rare
events with microscopic detail poses a ‘needle in a haystack’ problem. New approaches will need
to be developed to handle and analyze the large quantities of data that would be produced while
monitoring polymers for the occurrence of rare events. On the theory and computational side,
efforts to develop predictive capabilities will require improved force fields for more accurate
simulation of polymers and the availability of high quality thermophysical data against which
comparisons can be made.

4.3.2 National user facilities


User facilities at National Laboratories, NIST, and at university research centers play an
invaluable role in advancing the frontier of polymer science. The historical role of neutron
scattering in elucidating the structure and thermodynamics of polymers highlights this point. The
availability and accessibility of these facilities must continue to improve to ensure progress. The

78
planned or ongoing development of new facilities such as vSANS at NIST, NSLS-II at Brookhaven
National Laboratory, a second target station and associated neutron scattering instruments at
ORNL, and upgrades to synchrotron light sources are important parts of such continual
improvement. The polymer community must be proactive in engaging their counterparts at
National Labs at all stages of the planning and development of new or upgraded facilities to ensure
that polymer-specific experimental and computational needs are properly addressed.
Practices which help facilitate both the experimental work of users (e.g., relevant sample
environments and sample preparation facilities) and their analysis and handling of data should be
encouraged. Further, analysis methods should keep pace with advancements coming from theory.
In some cases, theory to properly analyze data from national facilities is either not available or not
readily accessible. The development and dissemination of theory to interpret data and associated
data analysis tools that make use of current theory is therefore encouraged. National facilities have
significant experience in large scale data management, which can be leveraged in efforts to develop
repositories for property data.
This report highlights the need for characterization capabilities (e.g., super-resolution
microscopy, spectroscopic AFM imaging, and specialized TEM) which are often beyond the reach
of individual investigators and university research centers. National user facilities are well placed
to provide these important and, in many cases, emerging resources in midscale instrumentation
centers.

4.3.3 Education
Integrated measurement, analysis, and prediction are critical to advancing the frontier of
polymer science. Developing competence and, eventually, expertise in the associated advanced
methods requires considerable effort from researchers. The ability to sustain this effort is in
question, particularly at smaller research organizations, which may not possess a critical mass of
users. This presents a risk to the community in terms of ensuring a strong base of competent users
of the advanced characterization and computational tools that are required to sustain progress.
Improved training for researchers is needed to ensure they make the best use of available facilities.
Some national facilities offer ‘summer schools’ or short courses in support of users of their unique
experimental and computational resources. Likewise, some professional societies (e.g., APS-
DPOLY) offer short courses which periodically focus on topics directly related to measurement,
analysis, and prediction. A renewed commitment to and expansion of such efforts are
encouraged—the implementation of ‘summer schools’ and well-developed online resources for
important topics of interest plus predictability in the availability of such resources would
significantly catalyze progress in using advanced structure-property characterization techniques
and computational tools.

4.4 Recommendations
• Support the development of new tools and methods that will provide structure and
property characterization with the resolutions required to address specific grand
challenges in polymer science.
• Encourage greater use of emerging or existing advanced experimental tools and methods
to address open questions.
• Establish accessible and well-managed cyberinfrastructure that facilitates the storage,
organization, curation, and retrieval of experimental and computational data on polymer
properties.

79
• Provide mechanisms to intensify theory, simulation, and experiments that further first
principles prediction of polymer structure and properties.
• Establish regional and/or national centers that provide important midscale facilities to fill
the gap between university-level and national-lab-level resources.
• Provide consistent, high-quality, geographically-distributed user training in advanced
experimental and computational techniques.

80
SECTION 5: ADVANCING PERFORMANCE
Discussion Leader
Karen I. Winey, University of Pennsylvania

Group Members
Luis Campos, Columbia University
Venkat Ganesan, University of Texas at Austin
Ryan Hayward, University of Massachusetts, Amherst
Mary Ann Meador, NASA
Anthony J. Ryan, University of Sheffield
David Simmons, University of Akron
Sindee Simon, Texas Tech University
Edwin L. Thomas, Rice University
Carl Willis, Kraton Performance Polymers

Special Acknowledgements
Philip J. Griffin, University of Pennsylvania

5.1 Introduction
Polymers are uniquely positioned to meet materials challenges that demand currently
unrealized properties and performance. Polymers with advanced performance, either alone or in
combination with other materials, will improve or replace traditional engineering materials such
as metals and ceramics. This will lead to new and exciting opportunities for highly energy-efficient
transportation, advanced coatings for buildings and infrastructure, precise delivery of therapeutic
agents, responsive and adaptive materials, and more. Block polymers also continue to be central
to the evolution of nanometer-scale, self-assembled patterns fundamental to microelectronic
devices and to nanostructured templates for novel and futuristic metallic metamaterials.
Grand Challenge: Advancing polymer performance requires simultaneous control over
multiple orthogonal properties through rational material design,
often in combination with other system components; and through
robust, controllable processing methods.
This requirement demands a holistic understanding of structure-processing-property
relationships based on a combined experimental, theoretical, and simulations perspective. Meeting
this grand challenge will involve addressing the following problems: (1) harnessing process
methods to deliver improved properties; (2) manipulating interfaces and surfaces to combine
properties; (3) controlling mass, charge, and wave transport in polymers; (4) manipulating and
controlling polymer crystallization; (5) designing polymer materials that sense, respond, and adapt;
and (6) controlling the non-linear and non-equilibrium responses of polymeric materials.

5.2 Harnessing Process Methods to Deliver Improved Properties


The methods used to process polymer-based materials profoundly affect their properties,
offering both fundamental challenges and opportunities for improving performance. For example,

81
in microelectronics applications, where polymer deposition is followed by solvent and thermal
annealing, techniques have been developed to precisely control block copolymer structure from
the nanometer to the micrometer and even centimeter scales.[1] Maintaining momentum in
microeletronics will require yet greater degrees of precise control and perfection of morphology.
Thin films and membranes intended for packaging and separation applications, ranging from a few
centimeters wide to several meters wide and produced on continuous casting lines, will rely on
control of crystallization, phase separation, and associated structural orientation across a wide
range of length scales.
An interesting technology that is nearly ready for commercialization involves preparing
asymmetric, isoporous membranes (pore size ~ 10-100 nm) for water treatment applications and
various filtration technologies (including active biologicals and purification of pharmaceuticals).
Integrating self-assembling block copolymers with Non-solvent Induced Phase Separation (NIPS)
has produced asymmetric, porous membranes on a large scale. Traditional membranes prepared
by NIPS afford high rate of water transport due to asymmetric structure and porosity, but they
suffer from a broad dispersity of pore sizes and low pore densities. This limits filtration utility; but
applying NIPS to an appropriately-defined block copolymer affords the self-assembled structure
shown in Figure 5.1. It has tightly-controlled, uniform pore size distributions and remarkably high
pore density. Both features are highly
prized in membrane technologies.
Another example where large-scale
processing of block copolymers
profoundly affects product performance is
the selectively sulfonated Nexar™
pentablock copolymers by Kraton
Figure 5.1. Cross-sectional (left) and top-view (right) SEM
images of an asymmetric isoporous membrane prepared by Performance Polymers. The nature of the
processing a block copolymer using NIPS techniques (scale casting solvent determines the selectivity
bars = 500 nm).[2] and permeability of the cast membranes.
The parameter space for polymer
processing is expansive and affects every class of polymer. Well-established methods (injection
molding, melt fiber spinning, spin coating, etc.), evolving methods (roll-to-roll processing,
electrospinning) and emerging methods (additive manufacturing, NIPS, spray coating) are largely
developed and refined empirically for each new polymer-based material. To deliver specific
combinations of properties, processing methods must be integrated at the earliest stages of design.
Strategies will require improved understanding of kinetics in complex polymer solutions,
development of advanced processing methods, and more data-driven methods that include
processing conditions. This is discussed in the next three subsections.

5.2.1 Kinetics in Complex Polymer Systems


The high viscosities and large interaction energies per chain typical of macromolecules often
produce structures and properties that are dictated by kinetic effects or processing method rather
than thermodynamics. Processing-defined structures include semi-crystalline polymers. The
mechanical properties of these materials depend sensitively on crystallization temperature, flow
conditions, and presence of nucleating agents, which affect the degree of crystallinity and the size
and shape of crystals and crystalline polymorphs.[3] Processing also influences the degree of
aggregation or dispersion of particulate fillers in composites.[4] It affects the size and connectivity
of phase-separated domains in polymer blends, dictating the mechanical toughness, electrical
conductivity, permeability, and optical characteristics. The fabrication methods for foams and

82
aerogels (which are kinetically trapped to have very high surface areas) are particularly sensitive
to solvent/polymer interactions, the stiffness of the polymer backbone, temperature, and many
other processing conditions. The glassy state is inherently non-equilibrium, yielding mechanical
properties that can vary dramatically from brittle to ductile depending on processing history and
aging during service life. Polymers containing various chemical moieties—including polar, ionic,
or acidic groups (which often result in self-association or selective interfacial interactions), present
added complexity.
Sensitivity to kinetic processes is an important asset to polymer-based materials and a major
challenge. It permits wide tuning of properties for a given set of raw materials, yet often presents
an immense design space defined by daunting arrays of possible processing variables. The field is
poised to make significant progress exploring this high-dimensional space rationally and
efficiently due to rapid improvements in theory and simulation tools plus development of high-
throughput characterization methods. Tightly integrating experiments with simulations and theory
is expected to substantially advance understanding of morphology development, phase transitions,
and property development during solvent removal, under applied forces, during temperature
changes, and through combinations of these variables. There are important and unresolved
fundamental scientific questions regarding phase transformations and diffusion in all-polymer
systems and in polymers combined with other materials, particularly nanoparticles. They will be
addressed simultaneously.

5.2.2 Advanced Processing Methods


Designing processes that optimally harness the intrinsic properties of polymers is a key
challenge. Understanding and controlling kinetically-defined structures and properties in complex
polymer systems addresses this challenge. This includes improving conventional processing (e.g.,
extrusion and film flowing), evolving techniques (e.g., roll-to-roll processing), and emerging
methods (e.g., additive manufacturing, electro-spinning). Opportunities to tune properties are
likely to be found in traditional polymer processing variables such as temperature, flow rate, and
solvent quality, but opportunities are also likely to be found in applying external fields (e.g.,
electric or magnetic). Engaging researchers from other disciplines such as electrical and
mechanical engineering to redesign processing equipment is expected to be fruitful. Often,
polymers that are present as minority components play critical and enabling roles, e.g., as
lubricants, dispersants, or thin coatings. In some uses, polymers are sacrificial elements that are
ultimately removed from the final product, e.g., in 3D printing, lithography, and templating of
mesoporous inorganic materials. The design of processes amenable to in situ, high temporal
resolution characterization methods that probe process variables, polymer structure, and properties
needs to be further developed. An important consideration that must be built into every advanced
process is the environmental footprint, including the toxicity or atmospheric activity of solvents;
energy demands; production waste; and the recyclability, reusability, or degradability of the final
product and any waste streams.
There is a need to develop tools to survey wide ranges of processing conditions to maximize
polymer properties. Ideally, such tools will permit optimization using small quantities of polymer
yet mimic large-scale continuous methods. For example, small scale extrusion devices have been
developed that provide access to manufacturing methods (such as multilayer film production) but
use minimal amounts of newly discovered experimental polymers. These techniques, combined
with strategies like statistical experimental design and analysis, allow reduction in the number of
experiments needed to evaluate a design space. They will lead to rapid and statistically relevant
improvements in desired properties and identify more efficient processing. These developments

83
will benefit from building relationships with industrial partners to gain greater understanding of
the realistic problems and needs in process scale-up. Engaging a variety of engineering disciplines
will advance high-throughput experiments pertinent to industrially-relevant polymer processing
methods.

5.2.3 Data-Based Methods that Include Processing Conditions


Applying data-based methods to predict polymer performance based on both structure and
processing has the potential to accelerate the development of new materials and products. Well-
defined and focused databases should be created with input from industrial partners to provide the
most important processing parameters and materials characteristics for predicting properties. A
variety of disciplines (operations research, statistics, computational science, and data sciences)
will be required to develop predictive tools specific to polymer processing and advanced
properties.
One computational approach being applied to complex problems in processing is inverse
design. Conventional computational approaches—and indeed many data-based methods—focus
on the forward problem of predicting properties based on specific aspects, such as the polymer
backbone chemistry, processing parameters, and thermal history. Since the design space is often
large with many possible variables, it is impossible to find true optimum solutions without luck or
leaps of intuition. Inverse design starts with the desired property and uses computational methods
to arrive at the global optimum. A few early examples of successful inverse design can be found
in the directed self-assembly of multi-block copolymers[5] or in the selection of optimal organic
photovoltaic polymers for improving the efficiency of solar cells.[6]

5.3 Control of Polymer Crystallization for Improved Performance


Whether considered by mass, volume, or dollars, a major fraction of commercially-marketed
polymers are semi-crystalline. Processing a semi-crystalline polymer involves shaping a liquid
through flow; using a die or a mold; then stabilizing that shape by a mechanism activated by either
a chemical reaction or the removal of heat or solvent. In some cases, the flow itself nucleates the
crystallization process. Achieving advanced properties in semi-crystalline polymers is inherently
complex. There are structural heterogeneities from 0.2 nm (e.g., monomer units) to 10 nm (e.g.,
random or extended polymer conformations) to 10 m (e.g., size and orientation of crystallites).
Manufacturers want to understand more about semi-crystalline polymers because they want to use
less material, lower energy costs, develop new combinations of properties from existing polymers,
and improve recyclability. Over the last decade, polymer crystallization has received limited
attention in academia, with the notable exception of time-resolved structural characterizations
using X-ray scattering at synchrotron sources in combination with rheology or processing.[7]
With growing computational power, simulations of structural evolution at the atomic-,
macromolecular-, and meso-scale have emerged. Theory and simulation are starting to provide
partial insights into how polymer architecture, molecular weight distribution, comonomers,
nucleating agents, and processing affect polymer crystallization, structure, and properties. Many
challenges remain, including the wide range of relevant length scales, local heterogeneities that
affect nucleation, and the non-equilibrium nature of most processing methods.[8] Thus, the ability
to select polymers and processing protocols a priori for an optimal combination of properties
remains limited. This challenge is growing as the variety of processing strategies expands from
conventional manufacturing methods (injection molding, extrusion and fiber and tape spinning) to
evolving (roll-to-roll processing) and emerging (additive) methods.

84
5.3.1 Bulk Polyolefins
The effects of heterogeneity on crystallization of polyolefins (mainly polyethylenes and
polypropylenes) and engineering polymers (like nylon and polyesters) are well known. Rather than
quiescent crystallization conditions, industry focuses on how typical industrial processes
(including extrusion, injection molding, and blow molding) impact crystallization and thereby the
final optical, mechanical, and thermal properties.
It is well established that flow-induced control of nucleation is crucial to the production of
many industrial products. Until recently, this understanding was entirely empirical, but using
model polymers with controlled architecture and molecular weight distributions has elucidated
some of the controlling molecular and flow parameters. But the field has not coalesced on a
common parameter set, nor has there been a clear and accepted theory amenable to prediction. In
essence, two parameters are responsible for the formation of shish‐kebab morphology by flow‐
induced crystallization: a minimum shear rate that is the inverse of the longest Rouse time of
molecules present in a polymer ensemble; and a specific work of flow (sometimes described in
terms of total strain) applied to the polymer. The magnitude of the specific work required to create
oriented (shish) nuclei, and thus the formation of shish‐kebab structures, depends on the polymer
relaxation times (and on chemical structure, molecular weight distribution, and flow geometry). It
has also been found that shish nuclei can be formed and exist at temperatures above the nominal
quiescent melting points. A schematic, correlations-based understanding of the process exists
(Figure 5.2), but the ab-initio quantitative link to morphology and properties remains elusive.
Studies of model systems (molecular weight, architecture, molecular weight distribution) are
required to make progress on the
theoretical front.
The underpinning science of
deformation processing in the solid state is
also well known. If the starting state is
well-characterized, and the stretching (in
one or two dimensions) well-defined, then
the orientation processes of spherulitic
structures can be modeled effectively. For
example, the new British £5 note is printed
on biaxially-oriented polypropylene,
where the phenomenology of the solid-
state processing into the structural
Figure 5.2. Schematic depiction of the formation of “shish- morphology and the resultant mechanical
kebabs” during flow-induced polymer crystallization. Stage properties are well understood (Figure 5.3).
1 involves stretching the longer polymers under flow that In addition to melt and solid state
upon collision produce stable point nuclei (Stage 2). Stage
processing, nucleating agents are routinely
3 involves aligning the point nuclei in rows and Stage 4 is
the transformation into fibrillar structures called shishes.[7] used to control (optical) clarity, cycle time,
and modulus. Nucleating agents are
typically small organic molecules added in minute amounts. They have been widely developed for
polypropylenes, polyamides, and polyesters because their slower intrinsic crystallization rate
allows greater control over property improvements. Fundamentals of polymer crystallization and
modeling kinetics during both isothermal and non-isothermal studies provide the necessary
framework for characterizing the effects of nucleating agents. While there is no method for a priori
prediction of a nucleating agent for a particular polymer, the common characteristics of the more

85
effective nucleating agents are well
established. Prior to crystallization, they
are typically insoluble in the melt, with a
higher Tm than the polymer, and are often
crystalline materials with structures similar
to the polymer of interest so as to promote
epitaxial crystallization on the nucleating
agent’s surface. A number of studies have
correlated the structure-property relations
for such specific nucleating agents, both
experimentally and through modeling.
The challenge in controlling
polyolefin properties remains to move
Figure 5.3. British £5 note printed on biaxially oriented beyond correlations and provide
polypropylene. From www.thenewfiver.co.uk. predictions. Filling the gap between
experiment, theory, and simulations will
take a massive interdisciplinary effort and a coherent, team-based approach. This will require
industrial scientists to identify the most pressing systems and conditions to study; synthetic
chemists to make well-defined polymers; and physical scientists to study the effects of polymer
architecture and flow on crystallization kinetics, morphology, and properties. These experimental
studies would naturally be guided by molecular rheology and dynamical theories and supported
by multi-length scale simulations. A more complete theoretical framework would accelerate the
development of better and more effective molecular design, processing methods, and nucleating
agents to expand semi-crystalline polyolefins for a wide range of innovative applications.

5.3.2 Other Bulk Polymers


Polymers possessing polar moieties with an ability to form intramolecular associations such
as H-bonding can dramatically alter the crystallization process. Typically, engineering polymers
such as polyamides and polyesters benefit from H-bond and dipole-dipole interactions that
dominate their crystallization behavior and subsequent mechanical properties. Higher-temperature
engineering polymers, such as liquid crystalline materials like Kevlar and the PEEK family, have
additional liquid-state preordering that enables and enhances specific interactions (like H-bonds
and stacking of aromatic groups) to increase association upon crystallization, and they are
amenable to flow enhancement. There are possibilities here for further advances, and the case of
step-growth polymers (like nylons and polyesters) can provide guidance for these advances. At
melt processing temperatures, these polymers are essentially dynamic due to trans-amidification
and trans-esterification, and the effective molecular weight is so low that entanglements—crucially
important in polyolefin processing—are not relevant. While this brings advantages, low viscosity
and low melt-elasticity tend to suppress flow-induced orientation.
Developments in X-ray synchrotron radiation sources and fast electronic area detectors have
allowed the investigation of polymeric materials at high temporal and spatial resolution. These
advances are best illustrated by studies of poly(ethylene terephthalate) (PET), a common polyester
and driven by the needs of the beverage industry: bottle weight can be minimized by taking
advantage of the crystallization and reorganization induced by strain in the (near) glassy state. The
fabrication of films and containers of PET typically involves mechanical deformation at elevated
temperatures close to its glass transition temperature, Tg. Such processing can have major effects
on the degree of polymer orientation and crystallinity and hence on the physical properties of the

86
material. Time-resolved X-ray diffraction techniques with millisecond time resolution have been
developed to investigate the strain-induced crystallization in PET under industrial processing
conditions. The micron-scale structural variation in orientation and crystallinity across the wall of
a PET soda-bottle can be measured using an X-ray beam as small as 2 µm diameter. These well-
established correlations between processing conditions and structural development are extended
to meet the necessary requirements for mechanical and permeability properties.
Supramolecular structures or protected intramolecular structures could be designed to take
particular advantage of flow-induced crystallization processing. Nature provides us with
inspiration in silk, an extrusion spun polymeric protein (or protein-carbohydrate alloy). Silk fibers
are well-known for their exceptional mechanical properties, achieved thanks to a unique
processing pathway that has evolved several times independently. It has been shown that silk is at
least 1000 times more efficient to process than a conventional synthetic polymer (HDPE) by
solidifying through dehydration as a result of flow. Much like an individual polymer chain in a
melt, a native protein and its closely-bound water molecules may be considered not as a solution
but as a single process-able entity: a nanocomposite termed an "aquamelt" that is split upon
exceeding an extremely low energetic processing threshold. To extend this to synthetic polymers
and supramolecular systems, it is necessary to fundamentally understand and then reverse engineer
these naturally-occurring systems and develop design rules for a new paradigm in polymer
processing. This will be achieved through a consortium of complimentary synthetic chemists,
structural biologists, and polymer rheologists, all supported by computer models.

5.3.3 Thin Films and Confined Systems


Thin and ultrathin films have many applications in microelectronics, sensors, coatings, and
solar cells. Much experimental effort has been made in recent years toward understanding the
crystallization and morphologies of thin films (thickness >10 Rg) and confined (ultrathin) films
(thickness ~ Rg) of both homopolymers and polymer blends.[9] For example, individual lamellar
growth in thin films is more complex than anticipated by bulk crystallization kinetics. Lamellae in
thin films do not grow at a fixed or uniform rate but rather in a seemingly stochastic manner,
becoming dormant at times, then starting to grow again. Furthermore, the stem length throughout
individual lamellae is variable rather than constant, as observed in bulk. These insights about
crystallization and morphology have been only loosely correlated to properties, so much work
remains.
The crystallization of ultrathin films depends strongly on the nature of the supporting
surface. A limited number of studies have probed crystallization as a function of polymer-substrate
interactions, and this effect is expected to be even more important in polymer blends when the two
polymers involved interact differently with the surface. There could be significant consequences
on growth rates and near-surface morphologies.
Confined polymer crystallization has been studied in droplets and nanostructures of other
topologies, especially in polymers infiltrated within inorganic templates, block copolymers, and
nanocomposites. The crystallization temperature generally decreases when a bulk material is
confined to the micro or nanoscale, where a fractionation between heterogeneously and
homogeneously crystallizing domains occurs. Polymers can also undergo surface nucleation at
large super-cooling or even true homogeneous nucleation inside these nano-sized domains.
Confinement often induces crystal orientation in ultrathin films and nano-layers. Preferential
orientation, both edge-on and flat-on, is common; and in cylindrical domains, both parallel and
perpendicular orientations are observed. The orientation is more evident when domain size is
reduced, but other variables (like molecular weight, affinity with the substrate and thermal history)

87
will determine the final orientation of the crystals. As broadly acknowledged in semi-crystalline
polymers, progress in confined systems is dominated by experimental studies of crystallization
kinetics and morphology development, while property measurements and theoretical and
simulations efforts are quite limited. To capture the full potential of semi-crystalline polymers,
comprehensive work is required to build both practical correlations and thorough scientific
understanding.

5.4 Manipulating Interfaces and Surfaces for Combinations of Properties


Commodity plastics are largely monoliths where the only interfaces are the surfaces. As
polymer-based materials have become more complex by introducing multiple polymers into a
single part (blends, multilayers, etc.) or by incorporating non-polymer materials (glass fiber
composites, nanocomposites, etc.), the importance of engineering the interfaces has grown. A
decade ago, it might have been acceptable to view the interfaces as a small region whose properties
are nominally the average of the individual material’s properties. However, recent experiments
challenge this concept by demonstrating that influence of polymer interfaces can be quite
significant. For example, perturbations of the glass transition can extend up to 350 nm from the
interface in an immiscible polymer blend, even when the interfacial thickness is only ~ 7 nm.[10]
Expanding the understanding of the underlying physics could revolutionize the development of
multifunctional, multicomponent materials.

5.4.1 Polymer-Polymer Interfaces in Polymer Mixtures and Block Copolymers


At the level of property determination, a significant number of open questions remain in
polymer-polymer interfaces. The polymer conformations near a polymer interface are different
from bulk, and the implications for interfacial properties are still unresolved. For instance, the
strength of polymer-polymer interfaces is critical for a variety of applications (polymer blends,
adhesives, coatings, multilayered packaging), and the paths to improving properties are often
limited to compatbilizing the interface. Moreover, for additive manufacturing, properties of
interlayer diffusion and adhesion are important and crucially depend on entanglement dynamics
in the interfacial regions.[11] The dynamics of entangled polymers at interfaces requires
transformational advances. Foundational questions remain, such as whether the length scale of the
interfacial properties is related to the width of the compositional transition. And while studies have
focused on the property perturbation normal to the interfaces, alterations in the chain conformation
will affect properties parallel to the interface, too. This only becomes more interesting at curved
interfaces.
Compared to immiscible and miscible polymer-polymer systems, microphase separated
block copolymers have considerably more interfacial area. In fact, the interfaces are so close
together, they are likely to impact one another. Perhaps the most striking observations have been
in the area of multifunctional materials, which draw upon the individual functional properties of
the two (or more) blocks, for instance, in mechanically strong ion conducting membranes or donor-
acceptor block copolymers. In addition to the role of morphology type and anisotropy, t reports
suggest the segmental and ion dynamics near and far from the microphase-separated interfaces are
distinct.[12] Enhanced transport properties have been achieved by manipulating the interfaces with
gradient copolymers, Figure 5.4.[13] The influence of these interfaces on the glass transition,
transport (electrons/holes, ions, small molecules), mechanical strength, etc., remains an open

88
question, and engineering the vast amount of interfacial areas in block copolymers could
substantially improve the properties for a variety of new applications.

Figure 5.4. Gradient copolymers comprising styrene and ethylene oxide exhibit enhanced Li-ion conduction
relative to their block counterpart.[13]

5.4.2 Polymer-Substrate and Polymer Nanocomposite Interfaces


Compared to polymer-polymer interfaces, far less is understood about the interfaces between
polymers and inorganic substrates. As a consequence of the intrinsically different chemical and
structural nature of the materials, even a quantification of the interfacial energy for different
combinations (and orientations) of materials remains challenging. Experiments have clearly
demonstrated the polymer-surface interactions can lead to properties for the polymer layer that are
markedly different from that of the bulk.[14] In almost all cases, the resulting property
characteristics depend sensitively on the polymer-substrate interactions. In some instances, they
exhibit length scales (~10 times the polymer radius of gyration) that are far larger than any quantity
physically rationalized based on conformations or density inhomogeneities. While some
experiments have been able to characterize some properties at a local level,[15] significant
opportunities remain in similar quantification of other properties.
Polymer nanocomposites are polymers modified by the presence of nanoparticles. For
inorganic nanoparticles, the unresolved issues discussed above are critically important due to the
exceptionally large quantity of interface in the nanocomposite. Complexities abound as the
influence of the nanoparticle chemistry, shape, size, size dispersity, and spatial distribution on the
polymer-nanoparticle interface is unknown; directly probing these curved interfaces remains an
experimental challenge. Predictive theoretical and simulation advances are hindered by lack of
systematic experimental knowledge: how does adding nanoparticles modify bulk thermodynamics
properties (equation of state, density, bulk modulus, thermal expansion coefficient, etc.)? Such
knowledge can provide crucial constraints on creating coarse-grained models with greater
material-specific predictive power, especially for dynamic-mechanical properties. As a
consequence, simple predictive rules to identify even whether a specified polymer/nanoparticle
combination is expected to be miscible (particles uniformly dispersed) or immiscible (particles
aggregated) have yet to be developed. The combination of van der Waals attractions between

89
particles and polymer-mediated depletion forces provides a strong thermodynamic driving force
for aggregation. While this issue has attracted considerable attention, more strategies are needed
to design chemical functionalization of nanoparticles and processing strategies, including external
fields, to produce optimal nanoparticle dispersions.
Interest in polymer nanocomposites has primarily arisen through property enhancements
observed (relative to the pristine matrix) in applications, such as mechanical strength, barrier
properties, electrical conductivity, fire retardancy, and optical properties. To fully advance
polymer nanocomposites, the origins of
these enhancements must be identified;
they are likely to differ by property. For
example, electrical conductivity in carbon
nanotube or silver nanowire
nanocomposites is dictated by the
percolated network and the junction
resistance between nanoparticles. For
mechanical properties, the
polymer/nanoparticle interfacial zone,
which can occupy a significant volume
fraction, is crucial; see Figure 5.5.
Clarifying how properties of interfacial
layers differ from bulk properties and the
length scale of such perturbations would
Figure 5.5. The high frequency modulus in polymer serve to significantly advance the ability to
nanocomposites is enhanced relative to the neat polymer improve polymer nanocomposites. This
due to the presence of mechanically reinforcing “bridges” challenge requires new experimental,
of interfacial polymer between nanoparticles (as depicted in
the AFM micrograph in the inset). Adapted from [16]. simulation, and theoretical methods that
thoroughly interrogate the interfaces.

5.5 Controlling Transport in Polymers to Expand Performance Portfolio


The transport properties of polymeric materials are enormously important to a wide array of
modern applications, including energy storage and conversion, water purification, food packaging,
optical coatings, gas separation, and thermal insulation. Polymeric materials are increasingly used
in these critical applications due to their low cost, light weight, flexibility, and processability. The
next central task is to harness these positive attributes and compound them to design new polymer-
based systems exhibiting well-defined transport properties that can be externally manipulated over
the course of their use.
These technological advances require an a priori understanding of how monomer chemical
structure, chain architecture, and mesoscale morphology collectively influence the transport of
mass, charge, and waves in polymers. This fundamental understanding requires comprehensive
advances in characterization techniques coupled with multiscale theoretical, simulation, and
inverse design methods. Advances in the ability to control the local chemical and physical
structures over a hierarchy of length scales promise to yield unprecedented regulation of the
transport behavior; new extreme values of transport; and maintaining or improving a variety of
orthogonal physical properties, such as mechanical strength, anti-fouling, and permselectivity.
Furthermore, polymers are being used more often as key components of complex material systems,
and it is therefore a crucial task of the polymer community to understand and control transport

90
across various intermaterial interfaces. A concerted effort to research these concepts has the clear
potential to dramatically expand the current portfolio of polymer applications in key technological
systems.

5.5.1 Transport of Charge Carriers and Low Molar Mass Molecules


Charged species (ions, protons, and electrons) and a range of small molecules (CO2,
methane, etc.) can be readily transported through polymers. The fundamental challenge with
regard to transport in polymers is developing design principles that produce materials with order-
of-magnitude increases in the permselectivity, whether pertaining to different charged species,
small solvent molecules versus ions, or varying chemistries of gas streams. This improvement in
permselectivity must simultaneously control the transport rates of the target species and maintain
other critical attributes of the material, such as mechanical integrity, chemical stability, and cost.
Advancing the ion transport in polymers could transform the global energy landscape due to
the critical role of membranes in rechargeable batteries, flow batteries, fuel cells, and
supercapacitors. In rechargeable batteries, solid polymer electrolytes are a potential solution to
formation of metal dendrites in high energy-density metal anode systems. They are a replacement
for the traditional and extremely flammable volatile solvents currently used in rechargeable
batteries. They also make possible thin, flexible, durable battery systems. One strategy to produce
high ion-conductivity/high-modulus solid-state polymer electrolytes includes lithium salt-doped
or ionic liquid-doped block polymers;[17] ion mobility is observed to be closely coupled to the
segmental diffusion process. More recently, membranes that exhibit polymer glass-like
mechanical properties, liquid-like ion mobility, and advantageous morphologies for ion
conduction have been reported.[18] This alternative strategy, inspired by the phenomenon of
“superionic” conductivity in solid state conductors, endeavors to design polymer materials in
which the ion transport process is decoupled from the characteristic polymer segmental transport
process and can become many orders of magnitude faster.[19] But significant work remains to
understand how to maximize this superionic transport mechanism.
Regarding the transport of electrons and holes in polymers, the anticipated applications are
organic electronics and photovoltaics. Specifically, controlling the transport of electrons and holes
in bulk heterojunction solar cells is critical to maximizing device efficiency.[20] In these systems,
short-lifetime excitons generated in the photoactive semiconducting polymer layer migrate over
distances of 1-10 nm to the large interfacial region between donor and acceptor polymer
nanophases and separate into free electrons and holes. If the electron and hole mobility in these
phases is vastly imbalanced, the interface becomes polarized and prevents the long-range
separation of charge. As such, new semiconducting polymer materials with tailored transport
properties must be developed to increase economic competitiveness. The efficient transport of
charge in these systems also crucially depends on the local morphology; optimally, a nanoscale bi-
continuous network of the donor and acceptor phases that creates effective charge transport
pathways to the current collectors while exhibiting characteristic length scales shorter than the
exciton mean-free-path. These concepts apply equally well to a broad range of semiconductor
applications, e.g., organic photodiodes, transistors, and thermoelectric materials.
Another realm ripe for advances in performance is chemical separation by the selective
permeability of small molecules. The growth of global greenhouse gas emissions from burning
fossil fuels requires new technologies for carbon capture and sequestration that are efficient,
economical, and long-lasting. Advances in design of polymer materials for CO2 sequestration
could dramatically mitigate the effects of emissions and provide a more carbon-neutral bridge fuel
to future sustainable and clean energy sources. Nanostructured polymer membranes with highly

91
controlled free volume holes (e.g.,
polytriptycenes, polynorbornenes) and
selective permeation networks offer a path
to enhanced separation efficiency
(especially CO2), expanding the empirical
limit of the Robeson upper bound into
currently uncharted territory (Figure
5.6).[21]

5.5.2 Transport of Waves


Elastic waves can transmit thermal
energy, mechanical energy, or sound, so
understanding how they interact with
polymers is important for designing next-
generation materials. Porous polymeric
materials are already widely used as high-
performance thermal insulators and sound
Figure 5.6. Robeson plot of CO2/CH4 selectivity vs. CO2
suppressors. Mechanical vibrations are
permeability for various classes of polymers. Newly
designed “thermally rearranged” (TR) polymers exhibit readily damped by solid polymers by
performance that vastly surpasses the conventional adjusting the distribution of molecular
Robeson upper bound.[21] relaxations to match the vibrational
frequency to be damped. Elastic waves can
be manipulated by designing materials with impedance-mismatched interfaces of specified size
and shape. Recent advances have been reported for highly non-classical behavior (e.g., phononic
band gaps and mechanical cloaking[22]) and for improved thermoelectric properties. Nano-
structured polymers, including polymer nanocomposites, provide new opportunities to control
orthogonal properties (e.g., smaller thermal conductivity and higher electrical conductivity).
Polymeric materials are also good for modulating electromagnetic waves. Periodic structures
of lengths from nm to nearly microns are possible with self-assembly of polymer colloids, block
copolymers, polymer nanocomposites, and photolithography. Polymers’ ability to change
dramatically their dimensions or optical or mechanical properties when the environment changes
makes for simple, versatile sensors (humidity, pH, temperature, mechanical forces, etc.). For
example, structural color has been dynamically-controlled using block polymers
swollen/deswollen with various solvents,
including ionic liquids.[23] The selective
addition of gain media (quantum
dots/rods/plates, perovskites, laser dyes)
into periodic polymer structures provides
further opportunities for novel applications
such as asymmetric transmission of light
signals (one-way light valves) (Figure 5.7).
The ability to tailor polymer
nanocomposites by encapsulating a wide
range of nanoparticles (different size,
Figure 5.7. Lamellar block polymers are selectively shape, density, stiffness, and
swollen with ionic liquid to form a tunable, soft photonic thermal/electrical conductivity) in a
crystal thin film coating. [23]
polymer matrix allows unprecedented

92
control of properties. For example, tailoring the chemistry, molecular size, and grafting density of
ligands attached to metallic nanoparticles means controlling local distances between nanoparticles
and tuning plasmonic resonances, leading to greatly enhanced electromagnetic fields in adjacent
regions and fine sensing of bound analytes that alter the local dielectric properties.

5.6 Environmental Stability, Instability, and Response


The versatility of polymers is clearly demonstrated by their wide range of responses to
stimuli. Some polymers are designed to persist for just a few hours when digested and other
polymers, like those used in construction, are expected to last for decades. There are new
opportunities for polymer-based materials that can shorten or extend product lifetimes and that
respond to a wider range of stimuli. Specific grand challenges include polymers with extended
stability under harsh conditions; polymers with engineered instabilities that trigger degradation
on-demand by various stimuli; and polymers that signal the presence of specific molecules or
environmental conditions.
Extreme environments generally cause unwanted instabilities in polymers. Vast
opportunities exist for new chemical and physical strategies to improve polymer stability for
medical applications. The gut, bloodstream, bone joints, and open wounds are examples of harsh
environments that change in response to disease and change the polymer itself. For the military,
lightweight plastics for ballistic protection remains a challenge. Perhaps the most challenging
opportunity for durable plastics is in space vehicles, withstanding vacuum but also the harsh
atmosphere of Mars and beyond. Ideal solutions are materials that are not only durable under harsh
conditions but go further and are able to adapt to their environment. Ideally, polymers and
composite materials could be designed with the ability to report when they are at risk of failure
and then either heal themselves or fail by design.
A grand challenge is to alter the fate of post-consumer commodity plastics.[24] Reprocessing
is widely effective, but impurities in the post-consumer waste streams prevent full recovery of the
pristine properties, which reduces the economic incentives for recycling. Modular compatibilizers
have the potential to decrease inhomogeneities within the impure recycled materials and may lead
to unprecedented properties. Block copolymers, branched macromolecules, and nanocomposites
are among the leading materials of choice in the efforts to meet these challenges. Recycling
polymers by chemical breakdown remains a challenge for the highly durable and most common
polymers produced worldwide: polyethylene, polypropylene, polystyrene, and poly(vinyl
chloride). New chemistries are required to
reduce polymers to well-defined
oligomers, functional derivatives, or even
their respective monomers. The monomers
will have obvious industrial uses, but new
technologies might be required to use the
other byproducts.
Being able to trigger degradation
could address the growing environmental
footprint of post-consumer waste streams
(Figure 5.8). Inefficient plastics recycling
Figure 5.8. Polystyrene ball accumulated at Midway
Atoll—a series of uninhabited islands in the middle of the has led to a surfeit of waste that endangers
Pacific Ocean—from plastic waste generated in Pacific Rim wildlife and ecosystems across the globe.
countries. Image reproduced from CNN. Here, innovative solutions are needed to

93
trigger degradation after the designated product lifetime. Currently, degrade-on-demand features
enable a variety of controlled-release technologies (e.g., chemical processing, agricultural,
personal care products), and further innovation is necessary to reduce plastic waste on land and
sea.
Responsive or smart polymers sense
and respond to their environment across a
wide range of stimuli (Figure 5.9). In
medicine, triggered degradation can be
beneficial for personalized drug delivery,
advanced imaging, and the combination of
diagnostics and therapeutics. In these
applications, the original polymer and the
subsequent degradation products must be Figure 5.9. Damage-reporting polymeric materials: only
the ruptured spheres luminesce under UV irradiation.[25]
free of adverse effects and be appropriately
secreted without causing damage—to the
patient or the environment. Responsive and degradable polymers have been exploited for 2D
patterning (e.g., photoresists), 3D scaffolds, or 3D templates that are subsequently used to fabricate
other materials into desired shapes (nanoscale to macroscopic). Additive manufacturing has
created new opportunities for these strategies. The automobile industry and advanced
transportation systems can benefit from self-healing coatings. They would minimize scratches and
dents and thereby extend product life. In agriculture, smart coatings might transform the industry
by reporting pathogen growth when crops are covered in bioplastic films for preservation. These
are a few examples of innovative and transformative areas of exploration in responsive and smart
materials.

5.7 Controlling Nonlinear and Non-Equilibrium Properties


The last 50 years have seen continuous progress towards a fundamental understanding of
polymer response to small perturbations from equilibrium, whether mechanical, optical, dielectric,
electronic, or acoustic. While much work remains, advances in understanding polymer properties
in this linear regime benefit from the existence of rigorous frameworks connecting macroscopic
linear regime response to equilibrium molecular structure and dynamics. In contrast, many of the
most critical performance properties of polymers reflect their response to large nonlinear
perturbations. Examples range from mechanical resilience or failure of an aircraft part to the
extreme limits of energy storage possible via electrically active polymers. However, the
understanding of polymer properties in the nonlinear regime, where external stimuli lead to large
deviations from equilibrium behavior, remains relatively rudimentary.
Establishing a fundamental understanding of polymer response to large (nonlinear) stimuli
is a grand frontier of polymer materials science and engineering. This gap between our
understanding of nonlinear response and its importance in real-world applications affects
numerous societal challenges. Examples include the need to engineer a new generation of
polymers and polymer-based materials that: 1) are mechanically resilient yet lighter weight for
aircraft and automobiles with improved fuel economy, 2) exhibit improved impact and blast
resistance and provide dramatically-enhanced protection for athletes, soldiers, and infrastructure,
and 3) possess optimized nonlinear electric and dielectric responses for the requirements of future
electronics, energy generation, and storage technologies.[26]

94
These examples highlight the critical importance of establishing a predictive understanding
of polymers’ nonlinear response to mechanical, optical, dielectric, electronic, and acoustic stimuli
as a function of chemistry, polymer architecture, processing, morphology, and thermal history.
Indeed, advances in the control of nonlinear response in polymers tend to have high-profile affects
on technology. For example, the last decade has seen transformational advances in the nonlinear
mechanical response of polymer hydrogels (see Figure 5.10). They are rooted in improved
chemical strategies for engineering toughness into networks from a molecular level. These
advances open the door to new life-altering advances in artificial tissues, ones that rival those in
our own bodies. This progress exemplifies the extraordinary opportunities and high potential for
transformational effects offered by new advances in the understanding and control of nonlinear
response in polymers.
Progress towards rationally
improving polymer nonlinear response has
been limited by a number of fundamental
challenges. Our current understanding is
based largely on empirical approaches. The
development of a rigorous, statistical
mechanical foundation for polymer
nonlinear properties could transform our
understanding of polymer properties,
leading to currently unrealized
advancements in performance. A second
challenge is the multiscale physics
involved in nonlinear response. For
Figure 5.10. Tough hydrogel exhibiting robust nonlinear example, the nonlinear mechanical failure
mechanical response at high (2,000%) strain.[27] of polymers via fracture can involve
physics spanning nine decades or more of
length scales. At the longest length scales, fracture involves large-scale material deformation (up
to meters or more), preceded by necking and shear-banding on the range of mm to cm. The fracture
event itself, though, requires bond breakage or chain pullout, events that occur on length scales of
~ 0.1 to 10 nm. Thus, nonlinear responses are intrinsically multiscale with nontrivial feedback
loops between behaviors spanning many orders of magnitude in both length and time. Overcoming
this challenge will require highly integrated experiment, simulation, and theory across the full
range of length and time scales. Of particular value would be new methodological advances in
integrating approaches at multiple scales, whether experimental or computational.
Coordinated theoretical, computational, and experimental efforts are needed to understand
nonlinear mechanical response in the relatively simple limit of amorphous homopolymers. There
are a number of areas in which a clear, mechanism-based understanding of homopolymer nonlinear
response appears to be within reach (if subject to a concerted decadal focus) and which would be
likely to have a large effect on polymer-based technologies. These include the following questions:
under specified large amplitude nonlinear deformation conditions, on what length scale are
polymer glasses effectively molten? What is the length scale of affine deformation? What are the
physics of non-entropic strain hardening, including what determines the relevant length and
modulus scales? What is the underlying physical nature of local plastic flow? What is the
underlying microscopic nature of nonlinear deformation in entangled glassy polymers, given that
entanglements and glassy networks store stress in different ways? How do these multi-scale

95
phenomena listed above ultimately determine mechanical instabilities such as fracture, shear
banding, crazing chain pull out, and the brittle-ductile transition? Answers to these and other
fundamental questions concerning the nonlinear mechanical response of homopolymers would
have a decisive effect on advancing the performance portfolio of polymeric materials.
Finally, we note that given the relatively immature understanding of nonlinear properties in
polymers, this is likely to be an area in which using formal inverse design methods will be
especially important. These methods have seen broad adoption in other materials communities in
designing, for example, the properties of metal alloys and the activity of biological
macromolecules. Early success has been demonstrated in the design of equilibrium structures and
properties of block copolymers.[28] Extending this to nonlinear responses in polymers could
revolutionize the control of end-use properties via chemical structure control and improved design
of processing operations. However, this will require overcoming a significant challenge: inverse
material design methods generally require an efficient and robust solution to the forward problem
of predicting, simulating, or measuring a material’s target properties. In the case of nonlinear
polymer properties, leverage of inverse design methods therefore requires high-throughput
methods for predicting or measuring these properties. A considerable focus should be placed on
developing improved simulation methods for predicting polymer nonlinear response efficiently
and with reasonable fidelity to experiment. Such an advance would enable more direct leverage of
inverse design methods, such as machine-learning and evolutionary algorithms, to advance the
understanding and design of polymer nonlinear response.

5.8 Recommendations
Ultimately, the commercial value of polymers and polymer-based materials is rooted in their
performance, and efforts to extend and diversify performance continue to produce improved
materials, new products, and innovative technologies. Key needs and recommendations include:

 Polymer processing frequently employs far-from-equilibrium conditions and non-linear


deformations, such that the design of processing methods to control polymer properties
is often accomplished by good intuition and empirical findings. There has been
considerable progress in correlating processing condition to polymer structure, most
notably in semi-crystalline polymers, but the connections to properties is less-well
developed. Efforts should be directed toward building experimentally-robust correlations
and predictive theories, simulations, and databases. Current and emerging processing
methods could then be manipulated to control multi-faceted polymer performance.
Particular opportunities exist for polymers containing nanoparticles, polymers that
possess distinct chemical moieties that produce nanoscale morphologies, and semi-
crystalline polymers.
 Polymers are nanoscale molecules and are often used under confinement, including thin
films and polymer nanocomposites. The interfaces and surfaces that impose this
confinement are known to influence properties and performance, although the
compositional profiles and the polymer conformations are better understood than the
effect on macroscopic properties. To fully optimize polymer performance, the role of
interfaces and surface in linear and non-linear properties must be developed.
 The importance of molecular, proton, ion, and wave transport through polymers has
grown in the last decade, as have the urgent need for improved membranes, polymer
electrolytes, advanced functional polymers, sound damping materials, etc. Further

96
development of polymers with tunable transport properties requires additional
understanding of the fundamental mechanisms of transport under various driving forces.
This would inspire new strategies for polymer design and processing methods that
capitalize on the fundamentals.
 The chemical diversity in polymers, like biopolymers, provides abundant opportunities
to engineer materials with an assortment of responses to stimuli, even multiple responses.
Technologies have yet to fully exploit the complete array of linear and non-linear
responses, chemically-triggered responses, reversible responses, and the absence of
response in extreme environments. To develop and advance such innovations, close
association with the envisioned applications will be advantageous.

97
SECTION 6: PARTNERSHIPS: ACADEMIA, INDUSTRY, AND
GOVERNMENT
Discussion Leaders
Pat Brant, ExxonMobil
Juan de Pablo, University of Chicago

Group Members
Shendra Baker, Synedgen
Kimberly Chaffin, Medtronic
Joe DeSimone, Carbon3D, University of North Carolina
Amalie Frischnecht, Sandia National Laboratory
Phil Hustad, Dow Chemical Company
Elsa Reichmanis, Georgia Institute of Technology
Matt Tirrell, University of Chicago

Special Acknowledgement
Joshua Lequieu, University of Chicago
Monirosadat Sadati, University of Chicago

6.1 Introduction
The US plastics industry accounts for over $500 billion in shipments of goods and is
projected to grow at nearly twice the rate of the GDP for the next few years (Figure 6.1)[1]. It is a
vital sector of the US economy; one of the few to have a trade surplus. However, maintaining
global competitiveness represents a continuous challenge—one that has grown in scope,
magnitude, and importance over the last decade. Accordingly, the US polymer industry has relied
on a relentless pace of technological innovation and will continue to rely on it. Many advances
behind that innovation can be traced to a productive partnership among academia, industry, and
government (AIG), which has been one of
the hallmarks of the US scientific polymer
enterprise.
For example, in polymer science and
engineering, the rapid growth of synthetic
polymers was driven by polymer synthesis,
engineering, physics, computing, and
manufacturing advances shared across
academic, industry, and government
laboratories. Today, lives are improved by
remarkable products ranging from
commodities (such as textiles and films) to
specialty materials (such as membranes,
Figure 6.1. Growth of US GDP compared to growth of medical devices, and implants (Figure
plastics, and rubber and related products. 6.2)).

98
Advances have unexpected benefits;
a scientific advance to solve a pressing
problem in one area often creates a solution
to a seemingly unrelated problem. Indeed,
the manifold innovations that arise from a
fundamental discovery are a hallmark of
polymer science and engineering. For
example, the polyimide denoted LaRC-SI,
originally developed for space
applications, was found to have improved
chemical resistance. It could better insulate
wires carrying electricity from a
pacemaker to the heart. This advance
meant longer lives and reduced healthcare
costs. This commercial success, as in Figure 6.2. Medical devices use many different polymers.
countless others, was aided by close Starting at the top and moving clockwise: Polyurethane
communication between industry and connector module on top of titanium can; ablation gun has
polycarbonate, ABS, polysulfone; diabetes pump has
academia, was made possible by polycarbonate housing; the AAA stent has a Dacron fabric
substantial federal support of fundamental mesh (courtesy Medtronic).
research, and was delivered through the
strong academic training of all participants. When the polyimide was inducted into the Space
Technology Hall of Fame in April, 2016 the press release stated: “When expertise is shared,
extraordinary results are possible.”[2]
Progress in polymer science and engineering has also benefited greatly from the stability—
and the constancy of purpose—of long-term collaborations among researchers in industry,
government, and academia. The societal benefits of these partnerships have been extraordinary yet
so numerous as to now seem obvious. Well-known examples include:

• Widespread replacement of metal and glass by plastic 1


• Hybrid packaging for food and beverages—a transformation that has reduced
transportation costs while delivering convenience, reduced food waste, and increased
safety.
• Water filtration with increasingly sophisticated polymeric membranes that allow access
to clean water. Entire countries depend on water filtration with polymeric membranes for
their survival.[3] The combined effect of innovations means millions more people will
have access to clean water.
• Light-weight structural polymeric materials have replaced heavier and costlier metallic
alternatives, leading to more efficient energy use, increased safety, and faster fabrication
cycles.
• Biocompatible and degradable natural polymer products have been purified and modified
from sustainable resources to provide innovation in packaging, waste disposal, and
medicine.

1This same success has resulted in critical assessments of sustainability (life cycle analysis), packaging reduction,
more aggressive recycling, and generally greater corporate, government, and public awareness of plastics and related
materials. These matters are further addressed elsewhere in this document.

99
Grand Challenge: Establish practical and efficient mechanisms that foster cooperation,
collaboration and working partnerships between industry, academia,
and government agencies and national laboratories for the purpose of
advancing basic research and technology development.
There are several ways strong partnerships are beneficial. For example, dynamic industries
have a never-ending supply of important problems driven by evolving market needs. Such
problems often require immediate or short-term attention attracting and absorbing considerable
human resources and capital. These factors can suppress and even stop industry’s focus on the
long-term, strategic research that would not only lead to more effective solutions but also to global
competitiveness and survival. It is not uncommon that behind the near-term solution is a
fundamental question or opportunity not yet addressed. If the biggest problems were clearly
communicated, they could influence academic and national lab research.
The federal government has long recognized the benefits of industrial-academic partnerships
and has been proactive, stimulating academic-industrial interactions and the translation of
academic discoveries into commercial opportunities. This is done through a variety of programs,
including, the Small Business Innovative Research (SBIR) program, Materials Research Science
and Engineering Centers (MRSECs), and Energy Frontier Research Centers (EFRCs). The Bayh-
Dole Act (1980) promotes transfer of intellectual property from federally-funded research to the
public. Without a doubt, it has also profoundly stimulated innovation.
National laboratories, such as the National Institute of Standards and Technology (NIST)
and the Department of Energy Laboratories, have aided key scientific and technological advances.
They have provided extraordinarily sophisticated, one-of-a-kind facilities (such as neutron sources
and synchrotrons) and the support staff necessary to facilitate collection and analysis of data.
Through a visionary initiative, Argonne, Oak Ridge, Lawrence Berkeley and Sandia National
Laboratories delivered strategically-important fabrication and characterization facilities and
services. Examples can be found at manufacturing.gov (National Advanced Manufacturing
Portal).
The pace and the very nature of change in polymer science and engineering has also
increased and will continue to do so. The design and adaptation of polymers into complex,
hierarchically-fabricated structures continues to accelerate. It is not uncommon, for example, for
polymeric films to consist of seven to fifteen layers of materials, where each layer could be a
composite or a blend of polymers, fabricated at speeds that rival or surpass that of newspaper
printing. The diverse array of skills required to advance the field of polymer science and
engineering demands greater collaboration across disciplines. That collaboration is essential if the
community is to keep pace with the new scientific challenges emerging with each technological
development.
We address measures to help foster industrial, academic, and governmental collaboration.
These measures will bring the diversity of thought and skills required to ensure the field of polymer
science and engineering continues to lead via productivity and innovation over the next decade.
More specifically, these recommendations are intended to create new intersections and improve
interfaces to increase the number of discoveries that reach commercialization and accelerate the
path to commercialization.

6.2 Over-Arching Opportunities.


Nearly one million people are employed by the polymer industry. Another 779,000 people
are employed in upstream industries that provide monomers to plastics manufacture1. Sustainable

100
and engineered polymers are growing and could expand these numbers. Importantly, employment
in plastics and specialty-polymer manufacture in the US has continued to grow while employment
in US manufacturing as a whole has shrunk. In all, the United States remains the largest polymer
producer in the world, with a substantial fraction of production exported to other countries. The
numbers above do not include the many positive downstream effects of plastics on the economy.
The future of the polymer industry appears bright, but this leadership position can only be sustained
through scientific and technological breakthroughs—advances fostered by a vibrant, fundamental
research environment. A changing landscape has posed important new obstacles that threaten
sustained development of the US polymer science and engineering enterprise. These obstacles are
outlined below.

6.2.1 Changing Landscape: Evolving Challenges and Opportunities at the Academia-


Industry-Government Interface
Multiple new factors affecting innovation have come into play since the previous report was
prepared a decade ago. The nature and magnitude of industrial research in the US has changed
significantly. Large segments of corporate polymer research have been cut or are migrating out of
the US. Other countries have sought to emulate the US research infrastructure and its inherent
formula for success so they can compete more effectively on a global scale. Highly educated,
skilled, and cost-effective work forces have been developed rapidly in Asia, Eastern Europe, and
South America. Moreover, local markets are emerging and growing to meet the local demands of
expanding middle classes. This has led to establishment of major industrial and government
research laboratories at locations that can better serve regional demands. China, Korea, Ireland,
and Israel are a few of the countries now attracting industrial research hubs and setting up new
government-funded labs. These changes inform industry’s perspective and its response to a
changing, increasingly complex and competitive global environment.
A new positive feature of these changes has been the eagerness of academics and industrial
entrepreneurs to develop startup companies. A common path has been to rely on university-based
discoveries, often generated with support from the federal government. Through private funding,
the startups develop the technology and intellectual property that must necessarily surround such
high-risk discoveries and make them attractive to larger-scale corporations. The advances are now
often aggregated for review in large innovation conferences. These startups provide a discovery
path for larger companies that lack the human and technological resources to pursue multiple
discoveries and high-risk engineering directions in parallel, especially without that the assurance
their investments will be profitable. In polymer science, such startups are found in the biomedical
industry, the electronics industry, the food industry, the additive manufacturing industry, and the
water filtration industry, to name a few.
Another new, positive change has been the rapid growth of shale gas production. This has
invigorated key segments of the US polymer industry by providing a source of low-cost feedstocks,
especially ethane, which is cracked into ethylene. This single development has spurred the
announcements of major integrated ethylene-polyethylene manufacturing projects and projects for
building blocks to other polymers. Including the ethylene-polyethylene projects, there are 264 in
all, now totaling over $164 billion in capital investment.[4] 2 They have just begun to generate new
jobs in monomer and polymer manufacture, projecting 738,000 jobs by 2023. These new jobs will
come from the integrated plastics industry manufacturers, fabricators, machine suppliers, and other

2 40% of the capital investment is complete or underway.

101
suppliers and from the wages spent by these workers.[5] Research into new routes for monomers
from shale gas fractions also has begun.
Polymers from renewable resources have progressed substantially since the last report.
These materials are particularly attractive for specialized applications; examples include high
optical-quality and scratch-resisting materials for lenses and windows. They rival and sometimes
surpass the attributes of established products, such as polycarbonates. Some of these new materials
are already being commercialized in international markets.
In manufacturing, extraordinary advances in additive processes will undoubtedly replace
established production methods, particularly for high-value-added parts or components. The
materials used in additive manufacturing were not originally intended for such processes;
opportunities abound for development of new monomers, chemical reactions, and polymer
composites as innovative additive processes are introduced.
Similarly, there is a continuing and even renewed appetite for research and development
focused on expanding the uses of polymers to meet increasingly demanding needs. These needs
also spur innovative manufacturing methods that can wed polymers with themselves and other
materials that break performance barriers. In sum, these advances combine to improve the quality
of life. While the breadth of envisaged applications and expected performance are daunting, the
record of innovation along the polymer delivery chain (Figure 6.3) has been remarkable. It bodes

Figure 6.3. (a) Illustration of the current polymer delivery chain and innovation scheme. (b) A future polymeric
materials delivery and innovation continuum scheme, where strong collaborations between universities, industries,
and government facilities aid the exchange of ideas through iterative, data-driven collaborations. (c) Successful
adaptation of changes to discovery-to-deployment for polymers can accelerate the number of new materials
reaching the market in an evolving US landscape.[5]

102
well for the future—provided industry and academia can rapidly adapt to the changing landscape.
They must facilitate collaborations and ideas, ease discussion and exchange along the discovery-
to-deployment delivery chain and innovation scheme.

6.2.2 Changing Attitudes Towards Polymeric Materials and Their Environmental Impact
Attitudes towards the sustainability and life cycle of materials, particularly polymeric
materials, have changed dramatically over the past decade. Major urban areas that did not have a
plastic recycling infrastructure now require plastic products be separated from trash. However,
recycling can be complicated by plastic articles comprised of materials that cannot be comingled.
These include different polymers; in some cases metal or metal-oxide layers found in high-barrier
applications. While recycling disposable plastic bags is not an issue, they have been banned for
other reasons.
To meet these growing societal challenges (including reduction of CO 2 emissions,
sustainability, healthcare, energy, and national security), the discovery of new polymeric materials
that address these concerns must be accelerated, as must their evolution to commercial use. Our
premise, underpinned by decades of experience, is that university-industry and government-
facilitated interfaces are key to enhancing the probability and the frequency of innovation.

6.2.3 A Renewed Case for Industrial-Academic Partnerships


Collaborations among researchers from different sectors lead to more intellectual diversity.
Researchers from academia, the national labs, and industry have different perspectives, tools, and
expertise. By working together, these researchers can contribute to a healthier scientific ecosystem
with an increased number of research ideas tested. This should lead to an increased number of
commercial successes (Figure 6.4).[6] Ultimately, universities educate the next generation of
inventors and industrial researchers. Exchange of ideas, needs, and concerns between academia
and industry is essential to maintain relevance and improve the value of undergraduate and
graduate education.
Many conditions have changed in development for polymer science and engineering and in
industry. Long-term corporate research has decreased substantially in the US. The disappearance
of large-scale corporate research
laboratories, especially in the last decade,
has created a void. At the same time,
startups providing market-driven
innovation are increasingly common.
International competition has never been
stronger and will continue to increase.
Polymers applications are increasingly
demanding. Polymers must be produced
under more stringent environmental
constraints. The case for facilitating
industrial-academic partnerships and
finding ways to overcome barriers to them
has never been stronger, but the
mechanisms to establish and maintain such
partnerships have remained fairly constant
Figure 6.4. Innovation success curve from concept to over the last few decades. Renewed
commercialization.[6]
engagement is needed between industry

103
and academia and that will require new strategies, new frameworks, and new government-
sponsored programs. In particular, the loss of corporate laboratories could be compensated for by
joint academic-federal research laboratories, if new mechanisms and interfaces are created. The
changing polymer research landscape has inadvertently created barriers—but also opportunities—
for the flow of people, ideas, and resources between different institutions and something about
removing those barriers and heightening the opportunities.
The recommendations below are largely a response to all of these changes. They address
intellectual property considerations; eroding perceptions about the merit of sustaining long-term
research horizons; the opportunities provided by the emergence of start-up companies and venture
capital; and mismatches between research areas considered important in industry versus those
considered important in academia or the national labs. We note other activities are also underway
that offer guidance for improving the academic-industry interface across a broader range of
disciplines.[7] Here, we identify nine actions—improvements or creations—that address the
changing R&D landscape, that should contribute to maintaining and strengthening AIG
interactions, and that should accelerate the rate of societally-effective innovation.

6.3 Recommendations
The following recommendations focus on the evolving landscape in polymer science and
engineering. Included are changes in organization, focus, complexity, competition, pace, and
ancillary technologies. These recommendations are primarily made to improve AIG lab interfaces,
thereby supporting and maintaining US leadership in polymer science and engineering and the
benefits to society that this leadership allows. However, the recommendations may relate to other
scientific and engineering research; thus, while the recommendations have been prepared
following an NSF-sponsored initiative, they should apply to other government agencies, too.

6.3.1 IP: Intellectual Property Ownership in Sponsored Research and Technology Transfer
Intellectual property was mentioned briefly on page 119 of the 2007 report, with the
recommendation, “Develop and support flexible IP practices that will allow timely translation…”
This takes on a heightened importance through the changes we have noted. We offer the following
concrete considerations and recommendations.
Industry funding of academic research has traditionally provided a valuable complement to
government support. But a variety of factors (including the increasing complexity of innovation
cycles and fewer or smaller corporate research labs) have weakened or fragmented efforts at the
industry-academic interface. At the same time, the relationship between academia and industry
continues to change; 3 a greater industrial focus on useful inventions from academic labs has led to
increased constraints and demands on intellectual property ownership.
While funding pioneering academic research is often desired, the complexity of the
negotiations around IP has led to large master agreements. Established industrial leaders select a
few universities to disperse funds for several research projects over a designated period.
Negotiations to initiate such corporate-sponsored research can take a year or more. In other cases,
negotiations are simply unsuccessful, and corporations look elsewhere for academic expertise.
Given that science capabilities in universities can contribute significantly to discovery and
development of commercial opportunities for companies and so sustain the growth of the US
polymer industry, it is essential to find solutions that reduce friction and overcome inertia at this

3For example, negotiations may now involve the effect of a collaboration on tax exempt bonds for buildings and
other infrastructure.

104
interface. To facilitate industry funding of academic research, we recommend NSF convene a
panel of NSF-funded scientists and of tech tnansfer experts (university and industrial). They will
assess best practices in sponsored research, arrive at recommendations, and provide a series of
reasonable IP template(s) to reduce the time and energy required to implement industrial-academic
research partnerships, thus increasing the probability of success.
By helping resolve IP issues noted above, there can be greater AIG collaboration, which can
implement the following two recommendations.

6.3.1.1 Collaborative Industry-Academic Research


With academic institutions showing increased interest in commercializing innovations,
collaborative relationships between industry, government, and academia ought to be mutually
beneficial. Entrepreneurship from faculty and students is also increasing, making small businesses
and technology transfer larger components of the industry-academic research program. Thus, in
the definition of industry, we include small, medium, and large companies
We recommend creating more effective communication between industry and academic
partners. Academics should tell industry about breakthroughs; industry should tell academia about
critical needs in fundamental polymer research. While industry-academic consortia are
opportunities for industry to present technical challenges to the academic community and can
enhance information transfer, mechanisms for actual collaborative research are critical to nurturing
these relationships and to ultimately turning communication into productive action.
The continued creation of and participation in opportunities for students and faculty to work
on projects of specific industry interest are important, for example, to work at industry sites; to
participate in small business innovation; and to share instrumentation, facilities, and databases. We
recommend revisiting and enhancing programs such as GOALI, I/UCRC, and RAPID to facilitate
and promote academic-industrial partnerships.
Key programs that have been successful include:

• I-CORPS: aimed at translating academic research into startup companies, public-


private partnerships that bring NSF researchers together with entrepreneurs, trainers,
and venture-capital specialists.
• I/UCRC: NSF research centers representing partnerships between industry,
governments, non-profits, and academic researchers to perform pre-competitive
fundamental research in areas of interest to US industry.
• STTR/SBIR: STTR is exclusively for tech transfer from universities to industry, a small
business interest. SBIR funds small businesses that typically are spawned from
academia but require a greater than 50% time commitment from the principal
investigator. While these are statutory mandated programs in which 11 agencies
participate, we recommend increased awareness and participation in the program for
those in academia who want to work with small business, start a new business or to
translate innovation to industry.

6.3.1.2 Cross-Functional Sabbaticals


To stimulate new ideas, establish new collaborations, and foster exchange of best practices,
we propose three actions:
1. Create a new sabbatical program, co-sponsored by NSF and other government
funding agencies, that will allow federally-supported researchers to spend extended
periods in industry or at a national lab. The program would let individuals expand

105
their career options by learning about a new field and familiarizing themselves with
emerging industrial/commercial/societal needs in polymer science and engineering.
This program could be extended to include postdoctoral fellows jointly-advised by
an academic and an industrial or government lab supervisor. Safeguards should ensure
postdoctoral fellows are engaged in ambitious research projects with clear goals and
not employed in day-to-day tasks best performed by regular company staff.
2. Rotate graduate students through industry labs; Through co-ops and internships,
undergraduate students have a chance to work in industry. Students with PhDs also
take post-doctoral positions in industry. Graduate students should also be able to work
in industry for an appropriate period .
3. Increase industry sabbaticals in academia. Industry sabbaticals are encouraged
through personal relationships and in some consortia (for example, the IPRIME
consortium at the University of Minnesota, where over 200 industrial researchers have
participated in its Industrial Fellows Program). Government does not fund such
sabbaticals. Nonetheless, NSF might take a more active role, encouraging these
sabbaticals to advance training, learn about a new field or an emerging one, and foster
more enduring collaborations.

6.3.2 Central Data Repository


There is value in sharing and standardizing data to promote technological advances.[8]
Mature disciplines, such as metallurgy, have created and standardized electronic databases,
thereby generating a resource contributing to improved steels and other materials. Polymer science
and engineering is a younger discipline with its own unique complexity. While data are available,
it is dispersed across sources, many of which are either not known or not readily accessible. These
data can be of variable quality and may have poor provenance. For theory and simulation-guided
design and to shorten the design and production cycle, all areas of the polymer science and
engineering community, from fundamental research to industrial products, would be well-served
by a central repository traceable data.
An initiative should be launched that will encourage polymer researchers to make a central
database with data about all polymers, polymer blends, and polymer composites. These data
should include details of the fabrication history and polymer state. Support for this could build on
efforts by European regulatory agencies and the FDA to make safety information about new
materials more readily available. The first step toward this goal could be a workshop/working
group, facilitated or sponsored by NSF or a group of agencies. This group would agree on standards
for experimental and simulation data on polymers. This effort would require full participation from
relevant government agencies, academia, and industry. Questions such as: “What kinds of data are
needed? In what format? How do we verify accuracy?” would be addressed to assess both value
and provenance for each entry.
Additionally, it would be valuable to make available old data from industry. Hopefully, these
data would no longer be proprietary, but the source of the data could be made anonymous, thereby
eliminating possible IP or legal considerations. To work with industry :

• a neutral third party could collect and protect, if necessary, industrial data
• machine learning could be used to mine the data and provide potential
suggestions/solutions to the user

106
The data would be available to everyone and accelerate understanding, discovery, and
technology development in polymer science and engineering. Such a repository would facilitate
insights in structure-property relationships and contribute to new discoveries in polymer science.
All researchers—industrial, academic, and government—would benefit from access to such data.

6.3.3 Industry-Academia-National Lab Summits


Research shows diversity of thought fosters innovation and helps drive market growth.[9]
As a scientific community, we understand the importance of diversity and we cultivate it. In the
case of polymer science and engineering, we require collaborations across disciplines to spark
practical use of our research. Established scientific meetings offer one critical mechanism to gain
insight from researchers in diverse fields. However, while these meetings are effective at
cultivating topical diversity, they have limited effectiveness building domain diversity, where
diversity is achieved when collaborations span the industry-academia-national lab domains. In
2013, American Chemical Society (ACS) published the historical demographics of meeting
attendees: 44% academic, 27% student, 16% industrial, and 5% national lab (8% did not
report).[10] 4 Given these statistics, it easy to understand why collaborations across domains might
only occur sporadically, through serendipity or through established relationships. Since diversity
drives innovation, why restrict innovation to topical diversity? By establishing a mechanism
through which key scientists can come together to discuss urgent matters and to build excitement
around the world envisioned if these matters had solutions, then we could collectively create
innovations driving fundamental and application-based research for decades.
Summits would allow scientists across domains to raise awareness and excitement around
research topics that have the scope and scale to affect the scientific community and the direction
and speed of innovation. Our proposal is as follows:

• Summits are hosted by NSF or a particular national lab.


• Establish a process to select summit topics, a chair for each, and a frequency and scale
for each meeting.
• Develop criteria for selecting government, industry, and academic participants for the
topic. This will assure participation across domains and take account of the changing
landscape in research and engineering.
• Set the goals and outputs for each summit (for example, review current challenges,
possible future applications, and latest research advances).

We anticipate such summits will foster collaborations and provide a tutorial on NSF awards
for such collaborative efforts.
NSF and other government entities recognize the importance of domain diversity and have
encouraged such diverse collaborations by establishing funding mechanisms specifically directed
at programs that incorporate Industrial-Academic-National Lab partnerships (IUCRC, GOALI
grants). But too often these funding mechanisms are under-used. We believe a primary reason for
the under-use is the inability to define mutually-beneficial research collaborations between these
diverse entities. Because they rarely find forums for interaction and scientific discussion, creating
these summits between them will deliberately foster discussion and exploration of research topics
critical to the world. They will fuel excitement around potential research topics and build the

4Conversely, many industrial researchers attend conferences, workshops, symposia and the like that are poorly
attended by academia.

107
foundation for diverse research collaborations, increasing the use of existing collaborative funding
tools. Research shows these domain-diverse research teams will create better outcomes and
potentially change the trajectory of innovation in our field for decades.

6.3.4 Facilitate Greater Use of Theory and Computation in Industry


Theory and computation are becoming less expensive, faster, and more effective in polymer
science and engineering. Modeling and theory can reduce the need for experiments, especially if
coupled with ready access to central data repositories and other initiatives aimed at accelerating
discovery through computation. This is valuable for multiple reasons, including: (1) providing
confidence in initially “noisy” experimental data; (2) identifying promising new materials via
predictive calculations; (3) guiding deployment of resources for key experiments; and (4) offering
an alternative to experiments where regulatory or safety concerns may preclude them initially. On
this fourth point, theory can guide safer laboratory and scale-up practices.
Industry generally does not have access to the theoretical and computational expertise
available at national labs and universities or their infrastructure. This is particularly true for small
companies. Even if industry has access, awareness is lacking, and it does not use the opportunity.
Our goal: create mechanisms to enhance industrial researchers’ awareness rof
theoretical/computational expertise and thereby increase use. Such mechanisms could be
branches of other initiatives already outlined. They could include support for postdoctoral scholars
supervised by academics at companies, support for industrial scientists on sabbatical in academia,
or academic researchers on sabbatical in industry.
To produce open source code readily used by both academic and industrial researchers, it is
essential to make investments supporting good software design and implementation and to support
the basic science. Often, the PI of a computational project may not have the expertise to include
best practices for state-of-the-art, scientific computing codes, such as automated testing and
complete documentation. NSF could support computation by including funding for software
development and other experts in grants to PIs developing new computational methods. We
recognize various branches throughout the NSF and DOE are responsible for development of
computational infrastructure, including software, but it is essential such branches get closely
involved with domain-specific scientists. Particularly, NSF and DOE should ensure proper
computational tools become available to the industrial and academic polymer communities.
Consistent with the ideas outlined below, national software centers should be established to
provide that knowledge at a broader scale. The resulting code, more stable and user-friendly, would
be much more usable by everyone, including industrial researchers.
There are national supercomputing centers that support high-performance computing. Such
centers are primarily devoted to using national massively parallel supercomputers. Parallelism is
effective for some types of calculations, but not others. As computing architectures continue to
evolve, we need regional centers that are nimble, that offer expertise in a wide array of computing
platforms, and that are scientifically broad to support the gamut of technological fields that
characterize industry. Some of the topics covered would range from multiscale modeling—starting
from electronic structure and classical atomistic modeling—and ranging all the way to finite
elements and other continuum approaches. Additional, essential topics include statistical analysis,
machine learning, data base construction, and the software necessary to pursue such endeavors.

6.3.5 Characterization and Small-Scale Manufacturing Facilities


With fewer corporate research laboratories, there is a need to expand characterization and
small-scale fabrication or advanced manufacturing facilities at academic institutions. These

108
regional facilities would support characterization needs, rapid prototyping of new materials, and
synthesis of polymers. Companies should be surveyed about the sorts of facilities they need, would
use, and would support, coupled with the scientific value of the facilities for academic and
government laboratory research. Facilities created will also need reasonable, low barriers for
industrial useand assurances they would serve a strategic purpose for broader, long-term societal
needs. Existing facilities, e.g., the Center for Functional Nanomaterials at Brookhaven National
Lab and the Carbon Fiber Technology Facility at Oak Ridge National Laboratory, could be models
for recommended centers.
We note there are already user facilities at national laboratories that provide some of these
services but not all. The model envisaged here is different. Barriers for immediate access would
be minimal; in addition to characterization, a center would provide synthetic equipment and
expertise; additionally, it would provide pilot-plant scale facilities to produce sufficient material
quantities for extensive testing and development. Such centers should be able to host industrial
and academic scientists for extended stays. They would engage multiple companies
simultaneously,and become a hub for industrial interactions, cross-fertilization, and growth. A
example of such a center is IMEC in Belgium, which accomplishes many of the above aims at a
truly international level.

6.3.6 Postdoctoral Fellowships and Engagement Programs for Startups


New programs at NSF provide support for startup companies. Examples include the STTR
and ICORPS initiatives, which have become an essential funding mechanism for small companies.
An addition to this ecosystem of support could be a postdoctoral program in which fellows would
be jointly supervised by an academic researcher and by scientists at a startup. A jointly-supervised
fellow could bring to bear the benefits of an academic environment and the expertise of an
academic advisor. The fellow could expose academic collaborators to emerging technologies in
the chosen field. The fellow would also have access to the support infrastructure provided by a
university, including characterization facilities. This form of collaboration would expose newly-
graduated researchers to an entrepreneurial environment from which they might benefit
considerably. The postdoctoral fellowship could be supplemented, if needed, with support for
access to university resources.

6.3.7 National Research Centers Devoted to the Advance of Polymer Science and
Engineering
Large, broad-spectrum research programs have been steadily declining nationally, both in
academia and at national laboratories. A simple overview of existing STCs, ERCs, and MRSECs
indicates the fraction of centers devoted to studying polymeric materials has steadily declined.
These centers have played an important role in fostering academic-industry partnerships, and their
decline at a time of growth for the polymer industry is unfortunate. An overview of EFRC centers
or the fraction of research devoted to the study of polymeric materials at national laboratories leads
to similar conclusions. Multiple research and development centers should be established, with
support from NSF and other federal agencies that provide support for continued innovation by the
US polymer industry. Such centers should offer synthesis, characterization, and advanced
computational modeling. Such centers should be embedded within universities, thereby providing
the highly-skilled technical expertise and state-of-the-art infrastructure necessary to compete in a
global industry. Through these centers, industry could influence and participate in research,
cooperatively conceived and executed with universities and national laboratories. Such centers
would also be national hubs for the exchange of ideas, for networking, and for innovation. A

109
natural outcome of such academic-industrial proximity is exposure of deep and interesting issues
in areas of polymer science (e.g., polyolefins, composites, etc.) that have gone out of fashion in
academia. For industry, these issues still merit serious fundamental attention and also address
important advances for society.

6.3.8 Industry Supported Polymer Science and Engineering Efforts in Academia


Most of the recommendations outlined above will undoubtedly represent additional burdens
and costs for federal funding agencies. The last decade has witnessed flat or declining funding for
polymer materials research; it is unrealistic to expect these burdens to be shouldered exclusively
by the federal government. It is essential for industry to assume some of the responsibility for
developing industrial-academic partnerships. They will not only lead to mutually-beneficial
advances but will also lead to development of a highly-skilled work force necessary to remain
competitive on a global scale.
This feature has been recognized by several major industrial players. For example, DOW
Chemical Company announced in 2012 a major investment in academic research, with a value in
excess of $125M over five years. Through their convening power, agencies such as NSF and NIST
could bring other industrial players to the table help organize investments, and even review far-
reaching projects involving this kind of academic-industrial partnership.

6.3.9 New Scalable Routes that Reduce the Cost of Monomers


Choosing one polymer over any other material for a given application is driven by the benefit
per unit cost. Thus, higher polymer cost can bar consideration in new applications, regardless of
utility. In turn, the polymer cost is governed substantially by the cost of the monomer(s) required.
This economic reality is also why polymers, in general, have been so successful. Because monomer
cost and availability is so central to opening new avenues for innovation and for pioneering
research to ignite that innovation, a new or greater focus should be established on monomer
feedstocks for polymers and polymeric materials. Finding scalable pathways that reduce
monomers’ cost and increase their availability will stimulate polymer science and engineering and
create new opportunities for industry.
Finally, when practical, NSF should initiate inclusion and collaboration with stakeholders
who want to improve AIG interfaces. These include professional organizations, such as the
American Chemical Society, and government agencies, such as NIST. In this way, these activities
sponsored and perhaps initiated by the NSF can be optimally leveraged.

110
SECTION 7: HIGHER EDUCATION AND DIVERSITY
Panel Discussion Leader
Jane Lipson, Dartmouth College

Special acknowledgement
Jeff Ting, University of Minnesota and University of Chicago

7.1 Introduction
A significant fraction of students graduating with a Bachelor's degree in chemistry or
chemical/materials engineering will end up working with polymers; but only a small percentage
of them will have had any education in the field. Chemistry and Physics departments typically do
not offer foundational courses with polymer-related content, let alone courses that focus on
polymers.[1] Materials Science and Engineering programs have hired more faculty with an
emphasis on polymers, and most curricula include some training in polymeric materials. However,
in many programs, traditional approaches to treating hard materials, like metals and ceramics, have
not been integrated with the fundamental principles governing polymers and soft materials. Even
in Chemical Engineering departments, a field that has embraced macromolecular science for
decades,[2] polymers and soft materials appear in core coursework only superficially.[3] With this
status quo in polymer engineering and science, it is a significant challenge to attract the students
needed to solve polymer-related problems of both fundamental and applied importance.
Generating interest among a diverse body of ambitious students is not enough; they must be given
the breadth of background needed to tackle demanding, multi-disciplinary problems. Indeed, the
polymeric nature and complex states of such materials requires researchers at all levels to integrate
computation, theory, and experiment to solve problems unique to macromolecular science and
engineering. Finally, having the best skill set is still only a start; the future scientific workforce
must also be independent and critical thinkers.
Grand Challenge: Design an education that embraces the breadth and complexity of
polymer science and engineering and to deliver it to a diverse student
body, making full use of technology to exploit expertise across the
field.

7.2 Challenges in education at the university level


To grow and maintain a truly diverse population of graduate students, postdoctoral
researchers, and career scientists, it will be crucial to increase the pool of undergraduates interested
in higher-level education. This will require recruiting from a significant breadth of colleges and
universities; growing the size of a nontraditional pool of graduate applicants may require
rethinking assumptions about competitive attributes. For example, the number of paid research
opportunities is limited, and many students—if not most—cannot afford to spend their summers
or leave terms doing unpaid research. On the other hand, all undergraduates have access to core
courses. A comprehensive background should include fundamental math and science and
introductory programming skills. This background should be as heavily weighted as research
experience when making decisions about graduate student offers.

111
A broader fundamental background may be available to all undergraduates, but completing
a materials science curriculum is far less common. Indeed, even having a dedicated materials
course as part of a physical sciences major is relatively rare. It is feasible to generate interest among
a broader swathe of students by developing material for introductory courses in chemistry, physics,
and engineering that incorporate one or more polymer-focused labs; use polymer-related examples
in lecture material; and explicitly connect the fundamentals of physics, chemistry, and engineering
to unique properties of polymeric materials. This linkage of properties and applications to
curricular fundamentals will only happen if relevant materials—lectures, problem examples,
foolproof labs, summaries related to applications—are made freely available to faculty and widely
advertised. A powerful approach would be to create a body that crossed professional societies and
brought together representatives from the main engineering, chemistry, and physics scientific
societies (e.g., the AIChE, the Society of Rheology, the ACS, the APS, the MRS). This idea and
other routes to curricular expansions are discussed below.
Another advance would be to create a polymer/materials major curriculum at institutions
where there is strong expertise but not enough faculty to cover the breadth of material required.
(Such a curriculum should span synthetic methods, characterization, and modeling.) In this
context, an online course component could enhance the undergraduate options. Introductory and
upper undergraduate level courses in materials science could be developed and recorded by top
faculty from many institutions. They would work together so no one department or faculty member
would face an onerous burden. This could be funded through a national agency call for proposals.
This call would challenge a collective of experts to work together, and the results could comprise
a significant piece of a scientist’s outreach effort for subsequent proposals. Courses developed via
the support of national science agencies and taught by acknowledged experts would enhance the
likelihood of degree credit.
Turning to graduate education, a workforce poised to make transformational advances in
polymer/material science and engineering will comprise experts in synthesis, characterization,
simulation, and theory who have the tools and vocabulary to work seamlessly across specialties.
Experimentalists should have the requisite skills for applying commonly-used theoretical methods
of analysis; should have acquired the ability to run simulation packages; and should be sufficiently
at ease with the vocabulary of theory and simulation to engage in productive discussions with
theorists and simulators. The gap between these groups must also be bridged by pragmatic
understanding from those on the theoretical/simulation side. These scientists should understand
the nature and availability of the sorts of experimental data relevant to the problems they are trying
to solve. They should also have a facility with using these data to test their models, and they should
aggressively communicate exactly how to apply their advances to analysis of experimental data.
Not all graduate programs have the local specialists needed to present the range of relevant
material in an effective way. The specialists in the most current experimental or computational
techniques could give targeted lectures; supplemental videos would demonstrate technical skills.
There could be a special YouTube or Vimeo station dedicated to these productions. There are other
routes: currently, there is a one-and-a-half day workshop in advance of the Polymer Physics
Division meeting at the American Physical Society March Meeting. The subject changes every
year; instead, there could be a rotating set of topics so graduate advisors could plan their students’
attendance. Additional workshops could be targeted to other meetings, such as those of the
Materials Research Society or the American Chemical Society. Summer school workshops could
enhance course curricula that had gaps, offering an intense, focused experience aimed at a subset
of core topics. Finally, perhaps there is potential for creating a summer ‘home’ for learning in

112
polymer materials science and engineering, along the lines of at least some of the educational
programs offered at the Woods Hole Oceanographic Institution. Indeed, Woods Hole[4] has
resources aimed at developing interest in oceanic sciences and engineering targeted at students
ranging from kindergarten through postdoctoral. It represents an aspirational model for the
polymer materials science and engineering community.
The above presupposes agreement about what constitutes a core curriculum; however, it is
clear the community would benefit from a broad discussion. The strategies outlined here will be
much easier to apply if there is consensus about fundamental background and the extent to which
material from biology, math, computer science, chemistry, physics, engineering (of different
types), and non-polymeric soft materials should be woven together. While respecting the need for
breadth in background, it will be important to maintain focus on the unique features and challenges
polymeric materials represent. Who would organize and moderate such a discussion? And who
would organize the creation and implementation of the agreed upon 'deliverables' (e.g., courses,
tutorials, workshops)? As suggested above, guaranteeing representation of all aspects of polymer
science and engineering (academia, national lab, and industry) would require a cross-society effort.
Funding could come from the societies from national labs, and from industry, since the
membership of these societies comes from all these areas.
These strategies should result in discussions of optimal content for a graduate curriculum in
polymer science and engineering. Cross-society efforts could lead the design of online specialized
courses for graduate students (and overview courses for undergraduates). Another graduate-level
need is making available expertise targeted to showing graduate students and postdocs how to use
new equipment and experimental techniques. On the modeling side, these web-based tutorials
could teach students how to run and modify simulation packages, but they could also cover
fundamentals of coding so interested students could apply the most current and sophisticated
modeling methods. These modules could be offered through scientific societies’ websites and
would allow students to access materials on specific skills at the time they are needed.
There is more to being educated then content alone. Industry needs PhDs who go beyond
being specialists; the most effective scientists are independent and creative problem-solvers. These
skills are grown. The process is best started before the graduate years, but there should be
conscious effort to strengthen those attributes in graduate school. A sense of the imperative to keep
learning—both information and skills—is critical, and it should be woven into the research and
curricular experience.
These remarks have focused on graduate student education, but they apply to the expanding
group of postdoctoral fellows who do research as part of their professional training. Many post
docs receive their first training in polymer science and engineering at this stage and access to
courses plays a critical role in empowering them before they enter the workforce.
Implicit and explicit bias in hiring and retention. Finally, it is useful to critique hiring and
retention choices, both in industry and academia. Diversity is not the enemy of meritocracy;
decisions are sometimes made using an all-too-familiar set of weighting factors that are not
unassailable (e.g., one being related to the names of academic institutions attended). In fact, such
metrics may be coded in a way participants do not fully appreciate. This is not a conversation that
happens easily, but it needs to happen, or the strategies above will not lead to real change.
A summary of many of the key issues and implications of bias can be found in a paper by
Handelsman and Sakraney for the White House Office of Science and Technology Policy.[5] This
article also references some sources that deal with tools for detecting and countering implicit bias.
Developing and maintaining a website providing resources for those involved in hiring and

113
retention decisions would provide a framework for understanding and applying best practices to
such decisions. Federal funding agencies could require all those involved in hiring decisions
participate in exercises designed to highlight implicit bias.[6, 7]
The most effective problem solving occurs when a group with a broad range of experiences,
viewpoints, and backgrounds works together. This speaks against the notion that hiring should
replicate the extant workforce. It will require innovation to achieve diversity, through both
outreach and education. By exploiting technological advances, an energetic and engaged scientific
cohort could make a substantial difference.

7.3 Recommendations
• Create a cross-disciplinary committee comprising representatives from appropriate
Divisions of scientific societies (e.g., the AIChE, the Society of Rheology, the ACS, the
APS, the MRS), including members from academia, industry, and the national labs. The
membership should also span areas of expertise, from synthesis to characterization to
both theoretical and simulation modeling. This group would: (1) coordinate efforts to
propose undergraduate and graduate curricula (see below); (2) outline resources required
for implementation of those curricula; (3) create a website to provide links for resources
already available; (4) annually assess which new efforts are successful; and (5) make
recommendations about fadeout of those that are not; (6) report to the appropriate
Divisions of the scientific societies that support this initiative.
• Fund a workshop (through the Societies and through NSF or other government agencies)
to develop an undergraduate and graduate curriculum for polymer/material science and
engineering. The cross-disciplinary committee described above could organize this
workshop and send representatives. Outcomes should include the proposed curriculum
and a description of required supporting resources (online and live) needed for academic
institutions to participate.
• Create a series of instructional modules aimed at graduate students and postdocs. The
modules would create new expertise in experimental techniques, simulation modeling
methods, and application of current theoretical approaches. The latter could be developed
using a simple programming tool, such as Mathematica. This could be co-funded by
government agencies and industry.
• Choose a subset of the cross-society polymer science and engineering group to assess
annually which efforts have been successful; to initiate fadeout for resources that did not
meet the need; and to update needed resources to keep them completely current. This
group would also make annual reports to the relevant divisions of the respective societies.
• Create a website that aggregates resources for best practices to counter implicit bias in
hiring and promotion. Attach a federal funding requirement that all who are responsible
for such decisions participate in exercises designed to highlight implicit bias.

114
SECTION 8: OUTREACH AND BROADENING PARTICIPATION
Panel Discussion Leader
Jane Lipson, Dartmouth College

Special Acknowledgement
Jeff Ting, University of Minnesota and University of Chicago

8.1 Introduction
Solving the most urgent problems of a rapidly-changing global landscape demands polymer
science and engineering workers who can adapt to new challenges. To realize the innovative
potential of this future workforce, and to have its work supported by the citizenry at large, they
must represent a diverse range of skill sets, perspectives, and backgrounds. Further, the polymer
community must become more effective at educating and engaging society about the important
role polymers play daily and polymers’ potential for resolving worldwide problems. If scientists
educate and engage, they will demonstrate the importance of polymer science advancement and
research, promote participation in science by a diverse community of citizens, and help build
informed goodwill towards national support for research.

8.2 Challenges and opportunities


We must work hard to strengthen the human infrastructure needed to accomplish our most
ambitious scientific goals. Waiting until college to reach out to students as they make career
choices is too late; interest in a scientific career must be sparked early. A student entering college
should already be reasonably well-prepared to take the core courses they need.[1] But that means
even very young children must be given opportunities to enter the world of scientific exploration.
Attracting a diverse body of ambitious students is not enough; they must be provided with the
breadth of background needed to tackle demanding, multi-disciplinary problems and to lead them
to independence and critical thought. We must pique the sympathetic interest of the much broader
segment of society—those who will not do scientific work but who will benefit from it. While the
challenges are perhaps discouragingly familiar, technological advances and a growing
entrepreneurial spirit point to new strategies. These strategies represent significant opportunities
to change the landscape of participation.

8.2.1 Creating a diverse body of student scientists


The makeup of the scientific community does not match that of the US population [2, 3] nor
even the profile of college students nationwide. Increasing the competition for the small portion
of the current population that is diverse is not a viable strategy; participation from a diverse
community must grow. How can we generate interest among such a broad section of the
population? Once engaged, how can we help these students move along a scholastic path that
demands the broadest possible education? This path begins in grade school and requires students
to build ever-deeper expertise. Without that expertise, students narrow their career options yearly
to the point that a career in science—or just a job—quickly becomes intellectually out of reach.
How can disadvantaged students be more competitive, particularly when they come from
institutions without the resources to produce competitive graduates? How can they go on to work
in industry or graduate school? This group needs to be able to move ahead in an unfamiliar

115
environment, with a cohort that does not share their challenges or background (e.g., financial
burdens, disabilities, etc.). The last question may be applicable at almost every level, from
elementary school to faculty level. The following section presents ideas on pre-university
educational mentorships, community outreach, and engagement opportunities to reach a diverse,
growing population that is trying to navigate terrain that might make individuals feel too much
like solitary explorers.
Grand Challenge: To "raise" a diverse community of future scientists and interested
citizens. This requires efforts that begin with young children and their
families, and continues through to launching ambitious high school
graduates towards university and beyond.

8.2.2 Outside the Classroom


Early developmental research in children has shown promising relationships between
explanation/exploration and fostering scientific reasoning.[4] Outside the classroom, there are
some places to reach out to young children and their parents to share the fun and excitement of
polymer science and engineering. Examples include shopping malls, farmer’s markets, other
seasonal events (e.g., autumn or winter fairs), and activities organized at local museums or public
places. These places can showcase the transformative benefits of polymers in everyday life under
place-related themes. For instance, at farmer’s markets, the idea of sustainable polymers derived
from renewable cellulose resources can be introduced. Indeed, along these lines, one could
emphasize that the display tables and essentially everything they contain are, themselves, largely
made of macromolecules in the form of biopolymers. Likewise, as a form of outreach, simple
science activities and demonstrations for the public could be carried out by middle and high school
students. A web search reveals multiple sites that promote polymer-related experiments and
science fair projects, but there is no source obviously associated with a national organization or
government agency that can serve as a resource for hands-on demonstrations. Resources for
outreach with a broader focus do exist; Table 1 highlights several examples from the American
Chemical Society, American Physical Society, and American Institute of Chemical Engineers. If
the subject material were carefully chosen, then this kind of participation could be credited as a
lab component of a science course and count for community service and leadership opportunities.
In addition, there are activities that exist but are not widely known. For example, Strange
Matter, a traveling materials science exhibition that is jointly funded by NSF, the Materials
Research Society (MRS), and industrial partners, can serve as a model for engagement.[5] As the
Strange Matter exhibition travels to different science centers across the country, local universities
and companies provide science projects, camps, and community projects tied to the materials
science. To date, Strange Matters has traveled to over 52 science centers with 5.3 million recorded
visitors. Furthermore, materials scientists have created a comprehensive collection of teacher
guides (linked with the national standard for various grades) and family guides (for in home
experiments). The guides have been translated into six different languages, and over 12 million

116
Table 1. National Organization Scientific Outreach for Pre-college Students.
Organization Program Description Link
American Science ACS Science Coaches are chemists (graduate https://www.acs.
Chemical Coaches students, professionals, or retirees) who volunteer org/content/acs/e
Society with an American Association of Chemistry n/education/outre
(ACS) Teachers (AACT) teacher member on an ach/science-
individual or group basis for one school year. coaches.html

National This program encourages chemists and chemistry https://www.acs.


Chemistry enthusiasts to increase awareness of chemistry at org/content/acs/e
Week the local level by hosting various chemistry- n/education/outre
inspired events throughout the country. ach/ncw.html

Kids & Kids & Chemistry is a community-based program https://www.acs.


Chemistry that connects scientists and children to do hands- org/content/acs/e
on science activities. Program volunteers include n/education/outre
ACS members, ACS Student Chapters, and ach/kidschemistr
corporate groups. y.html

American PhysicsCentral PhysicsCentral is an online platform that explains http://www.physi


Physical concepts in physics, scientists pursuing physics cscentral.com/
Society (APS) research, and how things in general work.

PhysicsQuest PhysicsQuest is a middle school competition based http://www.physi


on experiments centered on a mystery. Small cscentral.com/ex
groups or classes can register to receive free kits periment/physics
for this activity. quest/index.cfm

American K-12 Initiative This initiative uses the AIChE Volunteer Base to http://www.aiche
Institute of increase interest in science and engineering for K- .org/community/
Chemical 12 students by connecting individuals with k-12
Engineers program mentors and support.
(AIChE)

National Engineer Girl Introduces girls in grades 3-12 (with activities https://www.engi
Academy of broken into age groupings) to what Engineers do, neergirl.org/
Engineering and the steps in becoming an Engineer. Includes
interviews, activities, "ask an engineer", course
advice, scholarship information, and other
material.

people have downloaded them. Strange Matter has been self-sustaining since the first
NSF/industry/MRS investment, as the science centers pay a modest fee for bringing the exhibition
to the center. It could serve as a model for engaging communities.[5]
There are also strategies for drawing people to places they might not otherwise go (e.g.,
science museums) using social media and games. An unusual example of this involved a wildly
popular augmented reality game that made headlines in the summer of 2016, in part for drawing
unlikely museum goers into institutions across the United States.[6] To seem relevant to
generations to come, it will be key to engage people via the technology they are already using in
their daily lives, which means an online presence will be increasingly important.

117
Access to informal education on polymer science and engineering can be spearheaded on
the Internet. This online avenue can scale education availability and disseminate knowledge to
spark curiosity. Integrated video media platforms can also increase viewership and cultivate an
audience with scientific mindsets. Table 2 highlights channels on YouTube that regularly show
content dedicated to science, technology, and culture. These examples have millions of subscribers
and users in their communities, leading to broader interest in science and engineering and
dispelling common misconceptions. More formally, massive open online courses (MOOCs) have
emerged over the past decade, with some examples tabulated in Table 2. This route to reaching
large numbers of geographically-dispersed students has been implemented using a number of
strategies. Recently, Leontyev and Baranov have compared different MOOC approaches for
chemistry courses.[7]

8.2.3 Inside the Classroom from pre-K to Grade 12


Early developmental research has shown promising relationships between
explanation/exploration and fostering scientific reasoning.[4] For teachers, a number of resources
exist, depending on the location of the schools and on the funds available to take advantage of
science museum programs. Science Museums in St. Paul,[8] Pittsburgh,[9] Oklahoma City,[10]
Houston,[11] Boston,[12] and the Institution at Woods Hole[13] are just a few examples. While

118
there does not appear to be a comprehensive national listing of the major resources available to
teachers, some organizations compile listings of more-targeted resources. For example, the
National Education Association has a small set of links for teachers of preK-grade 12.[14] It is
clear STEM education in the classroom could benefit strongly from a site hosted by an agency
such as the NSF or perhaps through a joint effort of the National Academy of Sciences and the
National Academy of Engineering. Such a site would list programs by state and by city.
Information such as age range, level of cost, and possibility of sponsorship would make it much
easier for teachers across the country to judge what is truly available at each educational stage.
From grade school to high school, there is tremendous variation in what schools offer for
hands-on laboratory-type experiences; many schools do not have the financial flexibility or the
staff to engage students in hands-on labs. Some resources are available, but they require teachers
to learn the skills and background—either on their own, or through summer or other limited-time
courses—and then prepare, purchase, assemble, and run the labs. This is not unlike the hope that
people will turn to becoming fully-engaged home cooks by having recipes available (e.g., online)
and pointing people towards the nearest grocery store. In fact, numerous companies now ship
boxes with complete meal ingredients and preparation instructions—just open the box and begin.
This could be translated into a ‘Blue Labcoat’ experience: all precollege levels could have labs
that would be cheap, easy to run, and fit in a typical curriculum for the grade. Developing a series
of labs could be supported through proposals to national funding agencies. Industrial sponsors
could support the costs of putting together these ‘boxed labs’. The labs would be complete, such
that a teacher would only need to do the lab themselves, beforehand and then order a set. There
could be web resources—a video of someone doing the whole lab, additional information for
students who get interested/excited by the work, and so on. The goal would be to provide simple,
fun, creative labs that fit into the curriculum and that cost schools nothing but an application. A
similar concept could yield outreach-in-a-box to engage upper level students in outreach to
younger children, as mentioned above. Model polymer-focused outreach programs with simple
but fun experiments can be an engaging way to teach core polymer science concepts to students
and their local communities. [15]
One challenge to this goal is that there are no national curricular standards in science
education. In 2011, the National Research Council (NRC) of the National Academy of Sciences
(NAS) released “A Framework for K-12 Science Education”,[16] which has not been adopted at
the national level. Given the current status, a way forward could involve examination of the
education standards, state-by-state, to identify common requirements at different grade levels.
These could be summarized and published online, along with links to the standards in all fifty
states, to motivate the development of experiments that match curricular standards as broadly as
possible across the country.

8.2.4 Network of Support and Encouragement


Greater emphasis can be placed on early support and encouragement for students to become
involved in science and engineering. The approach could parallel elements of the culture around
youth sports, where young children interested in sports are encouraged to participate by families,
friends of parents, coaches, and mentors. Academic and industrial scientists can serve as scouts in
their own non-work environments, particularly those with school-age relatives. Local sections of
national societies could prepare information for their members about resources and opportunities
available in the community/region to get children engaged, lowering the barrier for its ‘member-
scouts’ to make concrete suggestions, and providing tools for ‘scouts’ to give both recognition and
encouragement.

119
Students who get interested in science need to know what will be required of them as they
advance. Students in economically-challenged regions are often unaware of the opportunities that
lie ahead. This is where the academy could play a significant outreach role, if it is generously
aimed. A web site could be cooperatively developed by individuals across a range of institutions
that would provide roadmaps for the needed accrual of skills, year-by-year, for a student to aim
towards a career in science. There could be links to study resources, including videos of students
from a wide range of backgrounds talking about their own experiences and challenges. There could
also be ‘success stories’—interviews with scientists working in a variety of fields and areas who
came from diverse backgrounds and were first in their families or neighborhoods to become
scientists.

8.3 Public Perception of Synthetic Polymers


Too often, “polymers” are synonymous with “plastics”, which evokes images of disposable
bags damaging the environment or compromised everyday consumer items, such as water bottles.
Given their ubiquitous appearance in everyday life, a relatively minor number of health and safety
issues are bound to arise and should not be discounted. However, as described in Section 1, these
giant molecules have driven tremendous advances in environmental and energy security, robust
infrastructure/communication, population sustainment, and health technology. The breadth of this
contribution and its tight linkage to the essential polymeric nature of the materials involved needs
to be much more effectively communicated. The message needs to be clear: polymers possess
properties that give them unique potential to address complex future challenges as societal needs
evolve.
This task will require strategic engagement from experts in academia, industry, and
government-supported entities, as discussed above. Motivated experts should engage the public at
local, state, and national levels. They should link synthetic polymeric materials to the myriad items
layered into all aspects of everyday life, from LEGO toys to artificial hips; from the skin (and
seats, and other components) of an airliner to Postit notes; from water purification filters to
computer chips and batteries.
An additional goal of these outreach efforts should be to position polymer science and
engineering as an endeavor that not only connects research in physics, chemistry, engineering,
mathematics, and biology, but also one that is critical to maintaining US leadership in technology
and innovation. This message should convey the excitement and potential of the field, along with
the need for future engagement of our brightest and most creative minds.
To accomplish the above, our community must aggressively share with the public some of
the energy and expertise we bring to our own work. We must reach out using all mechanisms:
social media, electronic and print news outlets, and personal engagement—wherever the public is
to be found.

8.4 Recommendations
• Develop a web site that aggregates resources for K-12 teachers in each state. Ideally,
there would be links to a state-by-state summary of core requirements at each grade level.
There would be links to outreach opportunities associated with science museums and
other science-related organizations in that state.
• Develop a web site that aggregates resources for teaching at all levels—K through post
graduate—offered through national organizations (such as the ACS and the APS) and
government-related something (such as the NSF or the NAS).

120
• Create an initial series of "Blue Labcoat" experiments that would meet core requirements
in all states. These could be generated on a one-by-one basis as an NSF requirement for
MRSEC-funded sites, for example, with one experiment to be developed over the life of
that grant. The requirement would also involve testing of the experiments in one or more
local schools.
• Develop partnerships with industry to fund distribution of "Blue Labcoat" experiments
in targeted school districts.
• Develop local and national programs aimed at cultivating visible proponents and
champions of polymer science and engineering. This would focus on imparting
techniques to improve scientific communication between scientists and engineers and to
the general public.

121
REFERENCES
Section 1
1. World population projected to reach 9.7 billion by 2050, 2015. United Nations
Department of Economic and Social Affairs.
http://www.un.org/en/development/desa/news/population/2015-report.html (December
15, 2016).
2. Handbook of Biopolymers and Biodegradable Plastics: Properties, Processing, and
Applications; Ebnesajjad, S; Elsevier/William Andrew: Amsterdam, 2013.
3. World Economic Forum, Ellen MacArthur Foundation and McKinsey and Company. The
New Plastics Economy–Rethinking the future of plastics; (2016)
http://ellenmacarthurfoundation.org/publications
4. Dittenber, D.B.; GangaRao, H.V.S. Critical review of recent publications on use of
natural composites in infrastructure. Composites: Part A. 2012, 43, 1419-1429. DOI:
10.1016/j.compositesa.2011.11.019
5. Impact of plastics packaging on life cycle energy consumption and greenhouse gas
emissions in the United States and Canada; Prepared for The American Chemistry
Council (ACC) and The Canadian Plastics Industry Association (CPIA); Franklin
Associates: 2014.
6. U.S. Energy Information Administration. International Energy Outlook 2016; Report
number DOE/EIA-0484(2016); May 11, 2016.
7. Williams, J.H., B. Haley, F. Kahrl, J. Moore, A.D. Jones, M.S. Torn, H. McJeon (2014).
Pathways to deep decarbonization in the United States. The U.S. report of the Deep
Decarbonization Pathways Project of the Sustainable Development Solutions Network
and the Institute for Sustainable Development and International Relations. Nov 25, 2014.
8. Wind Vision: A New Era for Wind Power in the United States; US Department of Energy,
Report DOE/GO-102015-4557, April 2015.
9. Balsara, N. P.; Newman, J. Comparing the Energy Content of Batteries, Fuels, and
Materials. J. Chem. Educ. 2013, 90 (4), 446–452. DOI: 10.1021/ed3004066.
10. Pesaran, A. A. Choices and Requirements of Batteries for EVs, HEVs, PHEVs
(Presentation). Present. CALSTART Webinar, 21 April 2011; Relat. Inf. NREL (National
Renew. Energy Lab). 2011.
11. Saha, P.; Datta, M. K.; Velikokhatnyi, O. I.; Manivannan, A.; Alman, D.; Kumta, P. N.
Rechargeable Magnesium Battery: Current Status and Key Challenges for the Future.
Prog. Mater. Sci. 2014, 66, pp 1–86. DOI: 10.1016/j.pmatsci.2014.04.001
12. Muskovich, M.; Bettinger, C.J. Biomaterials-Based Electronics: Polymers and Interfaces
for Biology and Medicine. Adv. Healthc. Mater. 2012, 3, 248-266. DOI:
10.1002/adhm.201200071
13. Yang, Z.; Zhang, J.; Kintner-Meyer, M. C. W.; Lu, X.; Choi, D.; Lemmon, J. P.; Liu, J.
Electrochemical Energy Storage for Green Grid. Chem. Rev. 2011, 111 (5), 3577–3613
DOI: 10.1021/cr100290v.
14. Darling, R. M.; Gallagher, K. G.; Kowalski, J. A.; Ha, S.; Brushett, F. R. Pathways to
Low-Cost Electrochemical Energy Storage: A Comparison of Aqueous and Nonaqueous
Flow Batteries. Energy Environ. Sci. 2014, 7 (11), 3459–3477 DOI:
10.1039/C4EE02158D.

122
15. Facchetti, A. Pi-Conjugated Polymers for Organic Electronics and Photovoltaic Cell
Applications. Chem. Mater. 2011, 23, 733-758. DOI: 10.1021/cm102419z
16. The Cement Sustainability Initiative: Our agenda for action.
http://www.wbcsdcement.org/index.php/about-csi/agenda-for-action (accessed December
15, 2016), World Business Council for Sustainable Development, page 20, published
June 1, 2002.
17. Conceptos Plasticos http://conceptosplasticos.com (accessed December 15, 2016).
18. VolkerWessels. http://en.volkerwessels.com (accessed December 15, 2016).
19. National Research Council, Division on Engineering and Physical Sciences; Committee
on Polymer Science and Engineering, Board on Chem. Sci. and Technology, Commission
on Physical Sciences, Mathematics, and Applications, Polymer Science and Engineering
The Shifting Research Frontiers; National Academy Press, Washington DC, 1994.
20. Sajjad, M.T.; Manousiadis, P.P.; Chun, H.; Vithanage, D.A.; Rajbhandari, S.;
Kanibolotsky, A.L.; Faulkner, G.; O’Brien, D.; Skabara, P.J.; Samuel, I.D.W.; Turnbull,
G.A. Novel Fast Color-Converter for Visible Light Communication Using a Blend of
Conjugated Polymers. ACS Photonics 2015, 2, 194-199. DOI: 10.1021/ph500451y
21. Eguza, S.; Wang, Z.; Chocat, N.; Ruff, Z.M.; Stolyarov, A.M.; Shemuly, D.; Sorin, F.;
Rakich, P.T.; Joannopoulus, J.D.; Fink, Y. Multimaterial piezoelectric fibres. Nat. Mater.
2010, 9, 643–648. DOI: 10.1038/NMAT2792
22. Roy, R.K.; Meszynska, A.; Laure, C.; Charles, L.; Verchin, C.; Lutz, J-F. Design and
synthesis of digitally encoded polymers that can be decoded and erased. Nat. Comm.
2015, 6, 7237. DOI: 10.1038/ncomms8237
23. Chemistry and Water: Challenges and Solutions in a Changing World, A White Paper
from the 6th Chem. Sci. and Society Symposium (CS3), Leipzig, Germany, September
2015. (http://www.rsc.org/globalassets/04-campaigning-outreach/policy/global-
challenges-policy/cs3-water-challenges-solutions-2016.pdf)
24. Shannon, M.A.; Bohn, P.W.; Elimelech, M.; Georgiadis, J.G.; Mariñas, B.J.; Mayes,
A.M. Science and technology for water purification in the coming decades. Nature. 2008,
452, 301-310. DOI: 10.1038/nature06599
25. Geise, G.M.; Lee, H-S.; Miller, D.J.; Freeman, B.D.; McGrath, J.E.; Paul, D.R. Water
Purification by Membranes: The Role of Polymer Science. J. Polym. Sci. Part B: Polym.
Phys. 2010, 48, 1685-1718. DOI: 10.1002/polb.22037
26. Los Angeles Times [Online], March 25, 2014.
http://articles.latimes.com/2014/mar/25/science/la-sci-sn-air-pollution-deaths-world-
health-organization-20140325 (accessed January 4, 2016).
27. Tullo, A.H. The Cost of Plastic Packaging. Chem. Eng. News, October 17, 2016, p 32-34.
28. Paper Pak Industries: The Natural Way to Extend Freshness.
http://paperpakindustries.com/products/ultrazap_xtendapak.php (accessed January 4,
2017).
29. Biomaterials: Important Areas for Future Investment, 2012 National Science Foundation
Biomaterials Workshop, Arlington, VA, June 19-20, 2012.
30. Smith KF, Goldberg M, Rosenthal S, Carlson L, Chen J, Chen C, Ramachandran S.
Global rise in human infectious disease outbreaks. J. R. Soc. Interface 2014, 11,
20140950. DOI: 10.1098/rsif.2014.0950

123
31. Welsh, A. What are the symptoms of Zika virus? CBS News [Online], August 2, 2016.
http://www.cbsnews.com/news/zika-virus-symptoms-how-do-you-know-if-youre-
infected/ (accessed December 19, 2016).
32. Yarsley, V. E.; Couzens, E. G. Plastics; Penguin Books: New York, 1941; pp 147–158.
33. Simon, D. T.; Gabrielsson, E. O.; Tybrandt, K.; Berggren, M. Organic Bioelelctronics:
Bridging the Signaling Gap between Biology and Technology. Chem. Rev., 2016, 116,
13009-13041. DOI: 10.1021/acs.chemrev.6b00146
34. Pew Research Center. Modern Immigration Wave Brings 59 Million to U.S., Driving
Population Growth and Change Through 2065; Washington, D.C., 2015;
http://www.pewhispanic.org/files/2015/09/2015-09-28_modern-immigration-
wave_REPORT.pdf (accessed Jan. 4, 2017).
Section 2
1. Deinssen, W., Winne, J. M., Du Prez, F. E. Vitrimers: permanent organic networks with
glass-like fluidity, Chemical Science 2016, 7, 30–38. DOI: 10.1039/c5sc02223a
2. Gentekos, D. T., Dupuis, L. N., Fors, B. P., Beyond Dispersity: Deterministic Control of
Polymer Molecular Weight Distribution, J. Am. Chem. Soc. 2016, 138, 1848–1851. DOI:
10.1021/jacs.5b13565
3. Matyjaszewski, K., Architecturally complex polymers with controlled heterogeneity,
Science 2011, 333, 1104–1105. DOI: 10.1126/science.1209660
4. Yokozawa, T., Yokoyama, A., Chain-growth condensation polymerization for the
synthesis of well-defined condensation polymers and pi-conjugated polymers, Chem.
Rev. 2009, 109, 5595–5619. DOI: 10.1021/cr900041c
5. Holm, M. S., Saravanamurugan, S., Taarning, E., Conversion of Sugars to Lactic Acid
Derivatives Using Heterogeneous Zeotype Catalysts, Science 2010, 328, 602–605. DOI:
10.1126/science.1183990
6. Smith, P. B., “Bio-Based Sources for Terephthalic Acid” Green Polymer Chemistry:
Biobased Materials and Biocatalysis, ACS Symposium Series 2015, 1192, Ch. 27, 453–
469.
7. Whited, G. M., Feher, F. J., Benko, D. A., Cervin, M. A., Chotani, G. K., McAuliffe, J.
C., LaDuca, R. J., Ben-Shoshan, E. A., Sanford, K. J., Development of a gas-phase
bioprocess for isoprene-monomer production using metabolic pathway engineering, Ind.
Biotechnol. 2010, 6, 152–163.
http://online.liebertpub.com/doi/pdfplus/10.1089/cap.2009.0073
8. Darensbourg, D. J., Making Plastics from Carbon Dioxide:  Salen Metal Complexes as
Catalysts for the Production of Polycarbonates from Epoxides and CO 2 , Chem. Rev.
2007, 107, 2238–2410. DOI: 10.1021/cr068363q
9. Hong, M., Chen, E. Y.-X., Towards Truly Sustainable Polymers: Metal-Free Recyclable
Polyester from Biorenewable Non-Strained -Butyrolactone, Angew. Chem. Int. Ed., 2016,
55, 4188–4193. DOI: 10.1002/anie.201601092
10. Jonesa, G. O., Yuena, A., Wojteckia, R. J., Hedrick, J. L., García, J. M., Computational
and experimental investigations of one-step conversion of poly(carbonate)s into value-
added poly(aryl ether sulfone)s, Proc. Nat. Acad. Sci. 2016, 113, 7722–7726. DOI:
10.1073/pnas.1600924113
11. Jia, X., Qin, C., Friedberger, T., Efficient and selective degradation of polyethylenes into
liquid fuels and waxes under mild conditions, Sci. Adv. 2016, 2, 6. DOI:
10.1126/sciadv.1501591

124
12. Montarnal, D., Capelot, M., Tournilhac, F., Leibler, L., Silica-Like Malleable Materials
from Permanent Organic Networks, Science 2011, 334, 965–968. DOI:
10.1126/science.1212648
13. Palma, A., Pasquarello, A., Ciccotti, G., Car, R., Cu++ and Li+ interaction with
polyethylene oxide by ab initio molecular dynamics, J. Chem. Phys. 1998, 108, 9933–
9936. DOI: 10.1063/1.476432
14. Borodin, O., Polarizable Force Field Development and Molecular Dynamics Simulations
of Ionic Liquids, J. Phys. Chem. B 2009, 113, 11463–11478. DOI: 10.1021/jp905220k
15. Feng, S., Voth, G.A., Proton Solvation and Transport in Hydrated Nafion, J. Phys. Chem.
B 2011, 115, 5903–5912. DOI: 10.1021/jp2002194
16. McDaniel, J., Choi, E., Son, C-Y, Schmidt, J.R., Yethiraj, A., Conformational and
Dynamic Properties of Poly(ethylene oxide) in an Ionic Liquid: Development and
Implementation of a First-Principles Force Field, J. Phys. Chem. B 2016, 120, 231–243.
DOI: 10.1021/acs.jpcb.5b10065
17. Fornace, M. E., Lee, J., Miyamoto, K., Manby, F. R., Miller III, T. F., Embedded mean-
field theory, J. Chem. Theory Comput. 2015, 11, 568–580. DOI: 10.1021/ct5011032
18. Kremer, K., Grest, G., Dynamics of entangled linear polymer melts: A molecular-
dynamics simulation, J. Chem. Phys. 1990, 92, 5057–5086. DOI: 10.1063/1.458541
19. Lee, J. Y., Shou, Z., Balazs, A. C., Modeling the Self-Assembly of
Copolymer/Nanoparticle Mixtures Confined Between Solid Surfaces, Phys. Rev. Letts.
2003, 91, 136103. DOI: 10.1103/PhysRevLett.91.136103
20. Qin, J., de Pablo, J. J., Ordering Transition in Salt-Doped Diblock Copolymers,
Macromolecules 2016, 49, 3630–3638. DOI: 10.1021/acs.macromol.5b02643
21. Maitra, A., Heuer, A., Cation transport in polymer electrolytes: A microscopic approach,
Phys. Rev. Letts. 2007, 98, 227802. DOI: 10.1103/PhysRevLett.98.227802
22. Webb, M. A., Savoie, B. M., Wang, Z., Miller III, T. F., Chemically specific dynamic
bond percolation model for ion transport in polymer electrolytes, Macromolecules 2015,
48, 7346–7358. DOI: 10.1021/acs.macromol.5b01437
23. Druger, S., Nitzan, A., Ratner, M., Dynamic bond percolation theory - a microscopic
model for diffusion in dynamically disordered-systems. 1. definition and one-dimensional
case, J. Chem. Phys. 1983, 79, 3133−3142. DOI: 10.1063/1.446144
24. Hibi, Y., Ouchi, M., Sawamoto, M., A strategy for sequence control in vinyl polymers
via iterative controlled radical cyclization, Nat. Comm. 2016, 7, 11064. DOI:
10.1038/ncomms11064
25. Canning, S. L., Smith, G. N., Armes, S. P., A Critical Appraisal of RAFT-Mediated
Polymerization-Induced Self-Assembly, Macromolecules 2016, 49, 1985–2001. DOI:
10.1021/acs.macromol.5b02602
26. Yu, Z., Tantakitti, F., Yu, T., Palmer, L. C., Schatz, G. C., Stupp, S. I., Simultaneous
covalent and noncovalent hybrid polymerizations, Science 2016, 351, 497–502. DOI:
10.1126/science.aad4091
27. Yashima, E., Ousaka, N., Taura, D., Shimomura, K., Ikai, T., Maeda, K., Supramolecular
Helical Systems: Helical Assemblies of Small Molecules, Foldamers, and Polymers with
Chiral Amplification and Their Functions, Chem. Rev. 2016, 116, 13752–13990. DOI:
10.1021/acs.chemrev.6b00354

125
28. Mavila, S., Eivgi, O., Berkovich, I., Lemcoff, N. G., Intramolecular Cross-Linking
Methodologies for the Synthesis of Polymer Nanoparticles, Chem. Rev. 2016, 116, 878–
961. DOI: 10.1021/acs.chemrev.5b00290
29. Erhardt, R., Böker, A., Zettl, H., Kaya, H., Pyckhout-Hintzen, W., Krausch, G., Abetz,
V., Müller, A. H. E., Janus Micelles, Macromolecules 2001, 34, 1069–1075. DOI:
10.1021/ma000670p
30. Altintas, O., Barner-Kowollik, C., Single-Chain Folding of Synthetic Polymers: A
Critical Update, Macromol. Rapid Comm. 2016, 37, 29–46. DOI:
10.1002/marc.201500547
31. Martinez, H., Ren, N., Matta, M. E., Hillmyer, M. A., Ring-Opening Metathesis
Polymerization of 8-Membered Cyclic Olefins. Polym. Chem. 2014, 5, 3507–3532. DOI:
10.1039/c3py01787g
32. Nakamura, A., Anselment, T. M. J., Claverie, J., Goodall, B., Jordan, R. F., Mecking, S.,
Rieger, B., Sen, A., van Leeuwen, P. W. N. M., Nozaki, K., Ortho-
Phosphinobenzenesulfonate: A Superb Ligand for Palladium-Catalyzed Coordination-
Insertion Copolymerization of Polar Vinyl Monomers. Acc. Chem. Res. 2013, 46, 1438–
1449. DOI: 10.1021/ar300256h
33. Lutz, J.-F., Ouchi, M., Liu, D. R., Sawamoto, M., Sequence-Controlled Polymers,
Science 2013, 341, 1238149. DOI: 10.1126/science.1238149
34. Rosales, A. M., Segalman, R. A., Zuckermann, R. N., Polypeptoids: a model system to
study the effect of monomer sequence on polymer properties and self-assembly, Soft
Matter 2013, 9, 8400–8414. DOI: 10.1039/C3SM51421H.
35. Barnes, J. C., Ehrlich, D. J. C., Leibfarth, F. A., Jiang, Y., Zhou, E., Jamison, T. F.,
Johnson, J. A., Iterative exponential growth of stereo- and sequence-controlled polymers,
Nat. Chem. 2015, 7, 810–815. DOI: 10.1038/NCHEM.2346
36. Pfeifer, S., Lutz, J.-F., A facile procedure for controlling monomer sequence distribution
in radical chain polymerizations.” J. Am. Chem. Soc. 2007, 129, 9542–9543. DOI:
10.1021/ja0717616
37. Li, J., Stayshich, R. M., Meyer, T. Y., Exploiting sequence to control the hydrolysis
behavior of biodegradable PLGA copolymers, J. Am. Chem. Soc. 2011, 133, 6910–6913.
DOI: 10.1021/ja200895s
Section 3
1. B. Lotz, S.Z.D. Cheng. A critical assessment of unbalanced surface stresses as the
mechanical origin of twisting and scrolling of polymer crystals. Polymer 46, (2005); 577-
610. DOI: 10.1016/j.polymer.2004.07.042.
2. G. Cesareni, M. Gimona, M. Sudol and M. Yaffe. Modular protein domains. (Wiley-
VCH, 2005).
3. H. An, J. Mike, K. A. Smith, L. Swank, Y. H. Lin, S. L. Pesek, R. Verduzco and J. L.
Lutkenhaus. Highly flexible self-assembled v2o5 cathodes enabled by conducting diblock
copolymers. Sci Rep-Uk 5, (2015). DOI:10.1038/srep14166.
4. S. Yang. Supramolecular lattices from tetrahedral nanobuilding blocks. Science 348,
(2015); 396-7. DOI:10.1126/science.aab0478.
5. M. Muthukumar and P. Welch. Modeling polymer crystallization from solutions.
Polymer 41, (2000); 8833-8837. DOI: 10.1016/S0032-3861(00)00226-3
6. P. Welch and M. Muthukumar. Molecular mechanisms of polymer crystallization from
solution. Phys. Rev. Lett. 87, (2001); 218302. DOI: 10.1103/PhysRevLett.87.218302

126
7. T. Yamamoto. Computer modeling of polymer crystallization—Toward computer-
assisted materials' design. Polymer, 50, (2009); 1975-1985. DOI:
10.1016/j.polymer.2009.02.038
8. X. He, J. Fan, J. Zou, K. L. Wooley. Reversible photo-patterning of soft conductive
materials via spatially-defined supramolecular assembly. Chem. Commun. 52, (2016);
8455-8458. DOI: 10.1039/c6cc03579e.
9. E. Gkikas, R. K. Avery, C. E. Mills, R. Nagarajan, E. Wilusz and B. D. Olsen. Hydrogels
that actuate selectively in response to organophosphates. Adv. Funct. Mater. 27, (2016)
1602784. DOI: 10.1002/adfm.201602784
10. B. M. Discher, Y. Y. Won, D. S. Ege, J. C. M. Lee, F. S. Bates, D. E. Discher and D. A.
Hammer. Polymersomes: Tough vesicles made from diblock copolymers. Science 284,
(1999); 1143-6. DOI: 10.1126/science.284.5417.1143.
11. M. Peterca, V. Percec, P. Leowanawat, and A. Bertin. Predicting the Size and Properties
of Dendrimersomes from the Lamellar Structure of Their Amphiphilic Janus Dendrimers.
J. Am. Chem. Soc., 133 (2011), 20507–20520. DOI: 10.1021/ja208762u.
12. C. S. Thomas, M. J. Glassman and B. D. Olsen. Solid-state nanostructured materials from
self-assembly of a globular protein-polymer diblock copolymer. ACS Nano 5, (2011);
5697-707. DOI:10.1021/nn2013673.
13. M. J. Glassman and B. D. Olsen. Arrested phase separation of elastin-like polypeptide
solutions yields stiff, thermoresponsive gels. Biomacromolecules 16, (2015); 3762-73.
DOI:10.1021/acs.biomac.5b01026.
14. R. M. Ho, M. C. Li, S. C. Lin, H. F. Wang, Y. D. Lee, H. Hasegawa and E. L. Thomas.
Transfer of chirality from molecule to phase in self-assembled chiral block copolymers.
J. Am. Chem. Soc. 134, (2012); 10974-86. DOI:10.1021/ja303513f.
15. H. K. Murnen, A. M. Rosales, J. N. Jaworsk, R. A. Segalman and R. N. Zuckermann.
Hierarchical self-assembly of a biomimetic diblock copolypeptoid into homochiral
superhelices. J. Am. Chem. Soc. 132, (2010); 16112-9. DOI:10.1021/ja106340f.
Section 4
1. Read, D.J., D. Auhl, C. Das, J. den Doelder, M. Kapnistos, I. Vittorias, and T.C.B.
McLeish. Linking Models of Polymerization and Dynamics to Predict Branched Polymer
Structure and Flow, Science, 2011. 333(6051): p. 1871-1874. DOI:
10.1126/science.1207060
2. Mike, J.F. and J.L. Lutkenhaus.Electrochemically Active Polymers for Electrochemical
Energy Storage: Opportunities and Challenges, ACS Macro Lett., 2013. 2(9): p. 839-844.
DOI: 10.1021/mz400329j
3. Liu, L. and H.J. Bakker.Infrared-Activated Proton Transfer in Aqueous Nafion Proton-
Exchange-Membrane Nanochannels, Phys. Rev. Lett., 2014. 112(25): p. 258301.
4. Roy, S., D. Skoff, D.V. Perroni, J. Mondal, A. Yethiraj, M.K. Mahanthappa, M.T. Zanni,
and J.L. Skinner.Water Dynamics in Gyroid Phases of Self-Assembled Gemini
Surfactants, J. Am. Chem. Soc., 2016. 138(8): p. 2472-2475. DOI: 10.1021/jacs.5b12370
5. Wang, C., D.H. Lee, A. Hexemer, M.I. Kim, W. Zhao, H. Hasegawa, H. Ade, and T.P.
Russell. Defining the Nanostructured Morphology of Triblock Copolymers Using
Resonant Soft X-ray Scattering, Nano Lett., 2011. 11(9): p. 3906-3911. DOI:
10.1021/nl2020526
6. Proetto, M.T., A.M. Rush, M.-P. Chien, P. Abellan Baeza, J.P. Patterson, M.P.
Thompson, N.H. Olson, C.E. Moore, A.L. Rheingold, C. Andolina, J. Millstone, S.B.

127
Howell, N.D. Browning, J.E. Evans, and N.C. Gianneschi. Dynamics of Soft
Nanomaterials Captured by Transmission Electron Microscopy in Liquid Water, J. Am.
Chem. Soc., 2014. 136(4): p. 1162-1165. DOI: 10.1021/ja408513m
7. Ullal, C.K., R. Schmidt, S.W. Hell, and A. Egner. Block Copolymer Nanostructures
Mapped by Far-Field Optics, Nano Lett., 2009. 9(6): p. 2497-2500. DOI:
10.1021/nl901378e
8. Boott, C.E., R.F. Laine, P. Mahou, J.R. Finnegan, E.M. Leitao, S.E.D. Webb, C.F.
Kaminski, and I. Manners. In Situ Visualization of Block Copolymer Self-Assembly in
Organic Media by Super-Resolution Fluorescence Microscopy, Chem. Eur. J., 2015.
21(51): p. 18539-18542. DOI: 10.1002/chem.201504100
9. King, J.T., C. Yu, W.L. Wilson, and S. Granick. Super-Resolution Study of Polymer
Mobility Fluctuations near c*, ACS Nano, 2014. 8(9): p. 8802-8809. DOI:
10.1021/nn502856t
10. McDaniel, J.G., E. Choi, C.-Y. Son, J.R. Schmidt, and A. Yethiraj. Conformational and
Dynamic Properties of Poly(ethylene oxide) in an Ionic Liquid: Development and
Implementation of a First-Principles Force Field, J. Phys. Chem. B, 2016. 120(1): p. 231-
243. DOI: 10.1021/acs.jpcb.5b10065
11. Segal-Peretz, T., J. Ren, S. Xiong, G. Khaira, A. Bowen, L.E. Ocola, R. Divan, M.
Doxastakis, N.J. Ferrier, J. de Pablo, and P.F. Nealey.Quantitative Three-Dimensional
Characterization of Block Copolymer Directed Self-Assembly on Combined Chemical
and Topographical Prepatterned Templates, ACS Nano, 2017. 11(2): p. 1307-1319. DOI:
10.1021/acsnano.6b05657
12. Arora, A., J. Qin, D.C. Morse, K.T. Delaney, G.H. Fredrickson, F.S. Bates, and K.D.
Dorfman. Broadly Accessible Self-Consistent Field Theory for Block Polymer Materials
Discovery, Macromolecules, 2016. 49(13): p. 4675-4690. DOI:
10.1021/acs.macromol.6b00107
13. Bates, F.S., M.A. Hillmyer, T.P. Lodge, C.M. Bates, K.T. Delaney, and G.H.
Fredrickson. Multiblock Polymers: Panacea or Pandora’s Box?, Science, 2012.
336(6080): p. 434-440. DOI: 10.1126/science.1215368
14. Chen, Q.P., J.D. Chu, R.F. DeJaco, T.P. Lodge, and J.I. Siepmann.Molecular Simulation
of Olefin Oligomer Blend Phase Behavior, Macromolecules, 2016. 49(10): p. 3975-3985.
DOI: 10.1021/acs.macromol.6b00394
15. Priestley, R.D., D. Cangialosi, and S. Napolitano.On the equivalence between the
thermodynamic and dynamic measurements of the glass transition in confined polymers,
J. Non-Cryst. Solids, 2015. 407: p. 288-295. DOI: 10.1016/j.jnoncrysol.2014.09.048
16. Pallon, L.K.H., F. Nilsson, S. Yu, D. Liu, A. Diaz, M. Holler, X.R. Chen, S. Gubanski,
M.S. Hedenqvist, R.T. Olsson, and U.W. Gedde.Three-Dimensional Nanometer Features
of Direct Current Electrical Trees in Low-Density Polyethylene, Nano Lett., 2017. 17(3):
p. 1402-1408. DOI: 10.1021/acs.nanolett.6b04303
17. Ducrot, E., Y. Chen, M. Bulters, R.P. Sijbesma, and C. Creton.Toughening Elastomers
with Sacrificial Bonds and Watching Them Break, Science, 2014. 344(6180): p. 186-189.
DOI: 10.1126/science.1248494
Section 5
1. Morris, M. Microelectron. Eng., 132, 207-217, 2015. DOI: 10.1016/j.mee.2014.08.009.
"Directed self-assembly of block copolymers for nanocircuitry fabrication."

128
2. Peinemann, K.-V.; Abetz, V.; Simon, P. F. Nat. Mater., 6, 992-996, 2007. "Asymmetric
superstructure formed in a block copolymer via phase separation."
3. Graham, R. S. Chem. Comm., 50, 3531-3545, 2014. DOI: 10.1039/c3cc49668f.
"Modelling flow-induced crystallisation in polymers."
4. Zeng, Q.; Yu, A.; Lu, G. Prog. Polym. Sci., 33, 191-269, 2008. DOI:
10.1016/j.progpolymsci.2007.09.002. "Multiscale modeling and simulation of polymer
nanocomposites."
5. Paradiso, S. P.; Delaney, K. T.; Fredrickson, G. H. ACS Macro Lett., 5, 972-976, 2016.
DOI: 10.1021/acsmacrolett.6b00494. "Swarm Intelligence Platform for Multiblock
Polymer Inverse Formulation Design."
6. O’Boyle, N. M.; Campbell, C. M.; Hutchison, G. R. J. Chem. Phys. C, 115, 16200-
16210, 2011. DOI: 10.1021/jp202765c. "Computational design and selection of optimal
organic photovoltaic materials."
7. Mykhaylyk, O. O.; Fernyhough, C. M.; Okura, M.; Fairclough, J. P. A.; Ryan, A. J.;
Graham, R. Eur. Polym. J., 47, 447-464, 2011. DOI: 10.1016/j.eurpolymj.2010.09.021.
"Monodisperse macromolecules–A stepping stone to understanding industrial polymers."
8. Wang, Z.; Ma, Z.; Li, L. Macromolecules, 49, 1505-1517, 2016. DOI:
10.1021/acs.macromol.5b02688. "Flow-Induced Crystallization of Polymers: Molecular
and Thermodynamic Considerations."
9. Prud’homme, R. E. Prog. Polym. Sci., 54, 214-231, 2016. DOI:
10.1016/j.progpolymsci.2015.11.001. "Crystallization and morphology of ultrathin films
of homopolymers and polymer blends."
10. Baglay, R. R.Roth, C. B. J. Chem. Phys., 143, 111101, 2015. DOI: 10.1063/1.4931403.
"Communication: Experimentally determined profile of local glass transition temperature
across a glassy-rubbery polymer interface with a T-g difference of 80 K."
11. Schnell, R.; Stamm, M.; Creton, C. Macromolecules, 31, 2284-2292, 1998. DOI:
10.1021/ma971020x. "Direct correlation between interfacial width and adhesion in glassy
polymers."
12. Singh, M.; Odusanya, O.; Wilmes, G. M.; Eitouni, H. B.; Gomez, E. D.; Patel, A. J.;
Chen, V. L.; Park, M. J.; Fragouli, P.; Iatrou, H. Macromolecules, 40, 4578-4585, 2007.
DOI: 10.1021/ma0629541. "Effect of molecular weight on the mechanical and electrical
properties of block copolymer electrolytes."
13. Kuan, W.-F.; Remy, R.; Mackay, M. E.; Epps III, T. H. RSC Adv., 5, 12597-12604, 2015.
DOI: 10.1039/c4ra15953e. "Controlled ionic conductivity via tapered block polymer
electrolytes."
14. Keddie, J. L.; Jones, R. A.; Cory, R. A. EPL, 27, 59, 1994. DOI: 10.1209/0295-
5075/27/1/011. "Size-dependent depression of the glass transition temperature in polymer
films."
15. Ellison, C. J.Torkelson, J. M. Nat. Mater., 2, 695-700, 2003. DOI: 10.1038/nmat980.
"The distribution of glass-transition temperatures in nanoscopically confined glass
formers."
16. Cheng, S.; Bocharova, V.; Belianinov, A.; Xiong, S.; Kisliuk, A.; Somnath, S.; Holt, A.
P.; Ovchinnikova, O. S.; Jesse, S.; Martin, H.; Etampawala, T.; Dadmun, M.; Sokolov, A.
P. Nano Lett., 16, 3630-3637, 2016. DOI: 10.1021/acs.nanolett.6b00766. "Unraveling the
Mechanism of Nanoscale Mechanical Reinforcement in Glassy Polymer
Nanocomposites."

129
17. Hallinan Jr, D. T.Balsara, N. P. Annu. Rev. Mater. Res., 43, 503-525, 2013. DOI:
10.1146/annurev-matsci-071312-121705. "Polymer electrolytes."
18. Schulze, M. W.; McIntosh, L. D.; Hillmyer, M. A.; Lodge, T. P. Nano Lett., 14, 122-126,
2013. "High-modulus, high-conductivity nanostructured polymer electrolyte membranes
via polymerization-induced phase separation."
19. Wang, Y.Sokolov, A. P. Curr. Opin. Chem. Eng., 7, 113-119, 2015. "Design of
superionic polymer electrolytes."
20. Lu, L.; Zheng, T.; Wu, Q.; Schneider, A. M.; Zhao, D.; Yu, L. Chem. Rev, 115, 12666-
12731, 2015. "Recent advances in bulk heterojunction polymer solar cells."
21. Sanders, D. F.; Smith, Z. P.; Guo, R.; Robeson, L. M.; McGrath, J. E.; Paul, D. R.;
Freeman, B. D. Polymer, 54, 4729-4761, 2013. "Energy-efficient polymeric gas
separation membranes for a sustainable future: a review."
22. Lee, J. H.; Singer, J. P.; Thomas, E. L. Adv. Mater., 24, 4782-4810, 2012. "Micro‐
/Nanostructured Mechanical Metamaterials."
23. Noro, A.; Tomita, Y.; Shinohara, Y.; Sageshima, Y.; Walish, J. J.; Matsushita, Y.;
Thomas, E. L. Macromolecules, 47, 4103-4109, 2014. "Photonic Block Copolymer Films
Swollen with an Ionic Liquid."
24. García, J. M. Chem, 1, 813-815, 2016. "Catalyst: Design Challenges for the Future of
Plastics Recycling."
25. Robb, M. J.; Li, W.; Gergely, R. C.; Matthews, C. C.; White, S. R.; Sottos, N. R.; Moore,
J. S. ACS Cent. Sci., 2, 598-603, 2016. "A Robust Damage-Reporting Strategy for
Polymeric Materials Enabled by Aggregation-Induced Emission."
26. Mecerreyes, D. Prog. Polym. Sci., 36, 1629-1648, 2011. "Polymeric ionic liquids:
Broadening the properties and applications of polyelectrolytes."
27. Sun, J.-Y.; Zhao, X.; Illeperuma, W. R.; Chaudhuri, O.; Oh, K. H.; Mooney, D. J.;
Vlassak, J. J.; Suo, Z. Nature, 489, 133-136, 2012. "Highly stretchable and tough
hydrogels."
28. Jaeger, H. M.de Pablo, J. J. APL Mater., 4, 053209, 2016. "Perspective: Evolutionary
design of granular media and block copolymer patterns."

Section 6
1. SPI, Size and Impact of the Plastics Industry on the US Economy. SPI: The Plastics
Industry Trade Association (2015), Washington, DC.
http://www.plasticsindustry.org/sites/plastics.dev/files/U.S.%20Size%20and%20Impact
%202015.pdf
2. http://www.medtronic.com/us-en/about/news/space-tech-hall-of-fame.html.
3. http://www.americanradioworks.org/segments/israel-water-technology-galilee/.
4. https://www.americanchemistry.com/Media/PressReleasesTranscripts/ACC-news-
releases/US-Chemical-Industry-Investment-Linked-to-Shale-Gas-Tops-164-Billion.html.
5. White, A. “The Materials Genome Initiative: One year on,” MRS Bull., 37(8), (2012)
715-716; DOI: 10.1557/mrs.2012.194 .
6. Stevens, G.; Burley, J. 3,000 raw ideas equal 1 commercial success! Research Technol.
Management 40(3) 1997, 16-27.
7. Mark. C. Cesa, Building strategic industry-university partnerships, Chem. Eng. News 94,
October 17, 2016, 38.

130
8. Fecher, B.; Friesike, S. and Hebing, M. “What Drives Academic Data Sharing?” PLOS
ONE, 10(2), (2015); DOI: 10.1371/journal.pone.0118053.
9. https://hbr.org/2013/12/how-diversity-can-drive-innovation
10. https://www.acs.org/content/dam/acsorg/meetings/nationalmeetings/why/convince-your-
boss-attendee-proposal-guide.pdf
Section 7
1. Kosbar, L. L.; Wenzel, T. J. J. Chem. Educ. 2017, Article ASAP. DOI:
10.1021/acs.jchemed.6b00922.
2. Tirrell, M. Chemical engineering of polymers: Production of flexible, functional
materials. Chem. Eng. Sci. 1995, 50, 4123–4141. DOI: 10.1016/0009-2509(95)00273-1
3. Cussler, E. L. The future of the lecture. AIChE J. 2015, 61, 1472–1477. DOI:
10.1002/aic.14807
4. Woods Hole Oceanographic Institution, http://www.whoi.edu/main/educate (accessed
Feb. 13, 2017).
5. Handelsman, J.; Sakraney, N. Implicit Bias. White House Office of Science and
Technology Policy.
https://obamawhitehouse.archives.gov/sites/default/files/microsites/ostp/bias_9-14-
15_final.pdf (accessed Feb. 13, 2017).
6. Lipson, J. E. G. A Simple Test Could Fix the Wage Gap Between Men and Women.
Quartz. April 9, 2013. https://qz.com/72058/a-simple-test-could-fix-the-wage-gap-for-
men-and-women/
7. Moss-Racusin, C. A.; Dovidio, J. F. Science Faculty's Subtle Gender Biases Favor Male
Students. Proc. Natl. Acad. Sci. USA 2012, 109, 16474–16479. DOI:
10.1073/pnas.1211286109
Section 8
1. Graham, M. J.; Frederick, J.; Byars-Winston, A.; Hunter, A. B.; Handelsman, J.
Increasing Persistence of College Students in STEM. Science, 2013, 341, 1455–1456.
DOI: 10.1126/science.1240487
2. Pew Research Center. Modern Immigration Wave Brings 59 Million to US, Driving
Population Growth and Change Through 2065; Washington, D.C., (2015).
3. National Science Foundation, "Women, Minorities, and Persons with Disabilities in
Science and Engineering" 2017, https://nsf.gov/statistics/2017/nsf17310/ (accessed Jan.
31, 2017).
4. Legare, C. H. The Contributions of Explanation and Exploration to Children's Scientific
Reasoning. Child Dev. Perspect. 2014, 8, 101–106. DOI: 10.1111/cdep.12070
5. Strange Matters, http://www.strangematterexhibit.com/ (accessed Nov. 5, 2016).
6. Sayej, N. "American art museums cautiously embrace Pokémon Go,"
https://www.theguardian.com/artanddesign/2016/jul/19/art-museums-pokemon-go
(accessed Nov. 29, 2016).
7. Leontyev, A.; Baranov, D. Massive Open Online Courses in Chemistry: A Comparative
Overview of Platforms and Features J. Chem. Educ. 2013, 90, 1533–1539. DOI:
10.1021/ed400283x
8. Science Museum of Minnesota, https://www.smm.org/educators/programs-your-school
(accessed on Feb. 13, 2017).
9. Carnegie Museums of Pittsburgh,
https://www.carnegiemuseums.org/interior.php?pageID=66 (accessed on Feb. 13, 2017).

131
10. Oklahoma Museum Network,
http://omn.sciencemuseumok.com/science_matters/outreach_programs (accessed on Feb.
17, 2017).
11. The Houston Museum of Natural Science,
http://www.hmns.org/education/educators/outreach-programs/ (accessed on Feb. 13,
2017).
12. Museum of Science, Boston, https://www.mos.org/traveling-programs (accessed on Feb.
13, 2017).
13. Woods Hole Oceanographic Institution, http://www.whoi.edu/main/educate (accessed
Nov. 29, 2016).
14. STEM Resources, National Education Association,
http://www.nea.org/tools/lessons/stem-resources.html (accessed Feb. 13, 2017).
15. Ting, J. M.; Ricarte, R. G.; Schneiderman, D. K.; Jiang, Y.; Saba, S. A.; Hillymer, M. A.;
Bates, F. S.; Reineke, T. M.; Macosko, C. W.; Lodge, T. P. Polymer Day: Outreach
Experiments for High School Students J. Chem. Educ. 2017, Article ASAP. DOI:
10.1021/acs.jchemed.6b00767
16. National Research Council. A Framework for K-12 Science Education: Practices,
Crosscutting Concepts, and Core Ideas; The National Academies Press: Washington, DC,
2012.

132
PERMISSIONS
Cover Art
Courtesy of Jeffrey M. Ting.
Report

Executive Summary
Figure: Courtesy of Frank S. Bates and Jeffrey M. Ting.

Section 1
Figure 1.1: Courtesy of Katie M. Greenman.
Figure 1.2: http://en.volkerwessels.com (accessed April 2017).
Figure 1.3: Courtesy of Ben Hsiao.
Figure 1.4: From Ref 26. http://articles.latimes.com/2014/mar/25/science/la-sci-sn-air-
pollution-deaths-world-health-organization-20140325 (accessed April 2017).
Figure 1.5: From Ref 28.
http://paperpakindustries.com/products/ultrazap_xtendapak.php (accessed April 2017).
Figure 1.6: From Ref 31. http://www.cbsnews.com/news/zika-virus-symptoms-how-do-
you-know-if-youre-infected/ (accessed April 2017).
Figure 1.7: Raw data from Ref 34. Courtesy of Jeffrey M. Ting.

Section 2
Figure 2.1: Reprinted with permission from Ref 2. Copyright 2016 American Chemical
Society.
Figure 2.2: From Ref 2. Courtesy of Kevin Noonan.
Figure 2.3: Reprinted with permission from Ref 9. Copyright 2016 John Wiley & Sons,
Inc.
Figure 2.4: From Ref 12. Reprinted with permission from AAAS.
Figure 2.5: Courtesy of Brett M. Savoie and Thomas F. Miller III.
Figure 2.6: Reprinted with permission from Ref 22. Copyright 2015 American Chemical
Society.
Figure 2.7: Courtesy of Barney Grubbs.
Figure 2.8A: Reprinted with permission from Ref 25. Copyright 2016 American
Chemical Society.
Figure 2.8B: From Ref 26. Reprinted with permission from AAAS.
Figure 2.8C: Reprinted with permission from Ref 27. Copyright 2016 American
Chemical Society.
Figure 2.8D: Reprinted with permission from Ref 28. Copyright 2016 American
Chemical Society.
Figure 2.8E: Reprinted with permission from Ref 29. Copyright 2001 American
Chemical Society.
Figure 2.8F: Reprinted with permission from Ref 30. Copyright 2016 John Wiley &
Sons, Inc.
Figure 2.9: Courtesy of Lisa Baugh.

133
Figure 2.10: Reprinted with permission from Ref 34. Copyright 2013 The Royal
Chemical Society.

Section 3
Figure 3.1: Courtesy of Stephen Cheng.
Figure 3.2: Reprinted with permission from Ref 3. Copyright 2015 Nature Publishing
Group.
Figure 3.3: From Ref 4. Reprinted with permission from AAAS. Courtesy of Felice
Macera and Shu Yang.
Figure 3.4A: Courtesy of Murugappan Muthukumar.
Figure 3.4B: Reprinted with permission from Ref 6. Copyright 2001 American Physical
Society.
Figure 3.4C: From Ref 7. Courtesy of Takashi Yamamoto.
Figure 3.5: Reprinted with permission from Ref 8. Copyright 2016 The Royal Chemical
Society.
Figure 3.6: Reprinted with permission from Ref 9. Copyright 2016 John Wiley and
Sons, Inc.
Figure 3.7: Reprinted with permission from Ref 11. Copyright 2011 American Chemical
Society.
Figure 3.8 (left): Reprinted with permission from Ref 12. Copyright 2011 American
Chemical Society.
Figure 3.8 (right): Reprinted with permission from Ref 13. Copyright 2015 American
Chemical Society.
Figure 3.9 (top): Reprinted with permission from Ref 14. Copyright 2012 American
Chemical Society.
Figure 3.9 (bottom): Reprinted with permission from Ref 15 Copyright 2010 American
Chemical Society.
Figure 3.10: Reprinted with permission from Ref 16. Copyright 2016 The Royal
Chemical Society.

Section 4
Figure 4.1: Courtesy of Composto, Caporizzo, and Goldman at the University of
Pennsylvania.
Figure 4.2: Courtesy of Molecular Vista, Inc. Available at http://molecularvista.com/
(accessed Mar. 2017).
Figure 4.3: Reprinted with permission from Ref 5. Copyright 2011 American Chemical
Society.
Figure 4.4 Top: Reprinted with permission from Ref 7. Copyright 2009 American
Chemical Society.
Figure 4.4 Right: Reprinted with permission from Ref 8. Copyright 2015 John Wiley &
Sons, Inc.
Figure 4.5: Reprinted with permission from Ref 11. Copyright 2017 American Chemical
Society.
Figure 4.6: Courtesy of Q. Chen, T. P. Lodge, and J. I. Siepmann.
Figure 4.7: Reprinted with permission from Ref 16. Copyright 2017 American Chemical
Society.

134
Figure 4.8: From Ref 17. Reprinted with permission from AAAS.

Section 5
Figure 5.1: Reprinted with permission from Ref 2. Copyright 2014 Nature Publishing
Group.
Figure 5.2: Adapted from Ref 7. Courtesy of Tony Ryan.
Figure 5.3: Image reproduced from www.thenewfiver.co.uk.
Figure 5.4: Reprinted with permission from Ref 13. Copyright 2015 The Royal
Chemical Society.
Figure 5.5: Reprinted with permission from Ref 16. Copyright 2016 American Chemical
Society.
Figure 5.6: Reprinted with permission from Ref 21. Courtesy of Benny Freeman.
Figure 5.7: Reprinted with permission from Ref 23. Copyright 2014 American Chemical
Society.
Figure 5.8: Image reproduced from
http://www.cnn.com/interactive/2016/12/world/midway-plastic-island/.
Figure 5.9: Reprinted with permission from Ref 25. Copyright 2016 American Chemical
Society.
Figure 5.10: Reprinted with permission from Ref 27. Copyright 2012 Nature Publishing
Group.

Section 6
Figure 6.1: Courtesy of Patrick Brant. Data plotted from Ref 1.
Figure 6.2: Courtesy of Medtronic.
Figure 6.3: Adapted from Ref 5. Courtesy of Jeffrey M. Ting.
Figure 6.4: Courtesy of Patrick Brant. Data plotted from Ref. 6.

135

You might also like