Compressing turbulence to improve inertial confinement fusion experiments (PRL)

Compression of a turbulent plasma. Credit: Seth Davidovits
Compression of a turbulent plasma. Credit: Seth Davidovits

By John Greenwald, Princeton Plasma Physics Laboratory

Physicists have long regarded plasma turbulence as unruly behavior that can limit the performance of fusion experiments. But new findings by researchers associated with the U.S. Department of Energy’s (DOE) Princeton Plasma Physics Laboratory (PPPL) and the Department of Astrophysical Sciences at Princeton University indicate that turbulent swirls of plasma could benefit one of the two major branches of such research.   The editors of Physical Review Letters highlighted these findings — a distinction given to one of every six papers per issue — when they published the results last week on March 11.

Lead author Seth Davidovits, a Princeton University graduate student, and Professor of Astrophysical Sciences Nathaniel Fisch, his thesis adviser and Associate Director for Academic Affairs at PPPL, produced the findings. They modeled the compression of fluid turbulence, showing effects that suggested a surprising positive impact of turbulence on inertial confinement fusion (ICF) experiments.

Stimulating this work were experiments conducted by Professor Yitzhak Maron at the Weizmann Institute of Science in Israel. Those experiments, on a Z-pinch inertial confinement machine, showed turbulence that contained a surprising amount of energy, which caught Fisch’s attention during a recent sabbatical at Weizmann.

In a Z-pinch and other inertial confinement (ICF) machines, plasma is compressed to create fusion energy. The method contrasts with the research done at PPPL and other laboratories, which controls plasma with magnetic fields and heats it to fusion temperatures in doughnut-shaped devices called tokamaks. The largest Z-pinch device in the United States is at the DOE’s Sandia National Laboratory. Other inertial confinement approaches are pursued at, among other places, the DOE’s Lawrence Livermore National Laboratory.

Present ICF approaches use compression to steadily heat the plasma. Methods range from squeezing plasma with magnetic fields at Sandia to firing lasers at capsules filled with plasma at Livermore’s National Ignition Facility. The presence of turbulence in the plasma is widely thought to increase the difficulty of achieving fusion.

But there could be advantages to turbulence if handled properly, the authors point out, since energy contained in turbulence does not radiate away. This compares with hotter plasmas in which heat radiates away quickly, making fusion harder to achieve. By storing the energy of the compression in turbulence rather than temperature, the authors suppress the energy lost to radiation during the compression.

The turbulent energy also does not immediately lead to fusion, which requires high temperature. This means a mechanism is needed to change the turbulence into the temperature required for fusion once the plasma has been compressed.

Davidovits used a software code called Dedalus to show that turbulent energy is increased during the compression, but then suddenly transformed into heat. As external forces in his simulation compress the turbulence to increase the energy stored within it, they also gradually raise the temperature and viscosity of the plasma. The viscosity, which describes how “thick” or resistant to flow a fluid is, acts to slow the turbulence and convert its energy to temperature. The viscosity started small so that the turbulence was initially unhindered. The rapid compression then kept the viscosity growing until it suddenly catalyzed the transfer of energy from the turbulence to the temperature.

In an experiment, this process would create the conditions for nuclear fusion in a plasma composed of the hydrogen isotopes deuterium and tritium. “This suggests a fundamentally different design for compression-based fusion experiments,” Davidovits said, “and a new paradigm for the inertial technique of producing fusion energy.”

He warns, however, that the simulation includes caveats that could diminish the findings. For example, the model doesn’t consider any possible interaction between the plasma and the containing capsule, and highly energetic turbulence might mix parts of the capsule into the plasma and contaminate the fusion fuel.

Nonetheless, the authors call the rapid transfer of turbulent energy into temperature during ICF experiments a “tantalizing” prospect that could benefit such research. And they note that their findings could lead to new understanding of the evolution of the relationship between the pressure, volume and temperature of a gas that is substantially turbulent. Determining this will be quite challenging, they say, “but the understanding will be important not only for the new fusion approach, but also for many situations involving the behavior of low viscosity compressible fluids and gases.”

This research was initiated through a grant by the Defense Threat Reduction Agency, a unit of the U.S. Department of Defense, and has been supported also by the DOE’s National Nuclear Security Administration through a consortium with Cornell University. Recently, the National Science Foundation and the Israel Binational Science Foundation combined funding opportunities to ensure further experiments at Weizmann on this topic and continued collaboration with the Princeton researchers.

PPPL, on Princeton University’s Forrestal Campus in Plainsboro, N.J., is devoted to creating new knowledge about the physics of plasmas — ultra-hot, charged gases — and to developing practical solutions for the creation of fusion energy. Results of PPPL research have ranged from a portable nuclear materials detector for anti-terrorist use to universally employed computer codes for analyzing and predicting the outcome of fusion experiments. The Laboratory is managed by the University for the U.S. Department of Energy’s Office of Science, which is the largest single supporter of basic research in the physical sciences in the United States, and is working to address some of the most pressing challenges of our time. For more information, please visit science.energy.gov.

Read the abstract or paper here.

S. Davidovits, N. Fisch, Sudden Viscous Dissipation of Compressing Turbulence. Phys. Rev. Lett. 116, 105004 – Published 11 March 2016.

This work was supported by DOE through Contracts No. DE-AC02-09CH1-1466 and NNSA 67350-9960 (Prime No. DOE DE-NA0001836), by DTRA HDTRA1-11-1-0037, and by NSF Contract No. PHY-1506122.

 

Down the rabbit hole: how electrons travel through exotic new material (Science)

Electrons on the surface of a Weyl semi-metal
Three-dimensional image using scanning tunneling electron microscopy of electrons on the surface of a Weyl semi-metal, a kind of crystal with unusual conducting and insulating properties. Image credit: Yazdani et al, Princeton University.

By Catherine Zandonella, Office of the Dean for Research

Researchers at Princeton University have observed a bizarre behavior in a strange new crystal that could hold the key for future electronic technologies. Unlike most materials in which electrons travel on the surface, in these new materials the electrons sink into the depths of the crystal through special conductive channels.

“It is like these electrons go down a rabbit hole and show up on the opposite surface,” said Ali Yazdani, the Class of 1909 Professor of Physics. “You don’t find anything else like this in other materials.”

The research was published in the journal Science.

Tantalum arsenide
Crystal structure of tantalum arsenide (TaAs), with Ta in blue and As in green. Image credit: Yazdani et al., Princeton University.

Yazdani and his colleagues discovered the odd behavior while studying electrons in a crystal made of layers of tantalum and arsenic. The material, called a Weyl semi-metal, behaves both like a metal, which conducts electrons, and an insulator, which blocks them. A better understanding of these and other “topological” materials someday could lead to new, faster electronic devices.

The team’s experimental results suggest that the surface electrons plunge into the crystal only when traveling at a certain speed and direction of travel called the Weyl momentum, said Yazdani. “It is as if you have an electron on one surface, and it is cruising along, and when it hits some special value of momentum, it sinks into the crystal and appears on the opposite surface,” he said.

These special values of momentum, also called Weyl points, can be thought of as portals where the electrons can depart from the surface and be conducted to the opposing surface. The theory predicts that the points come in pairs, so that a departing electron will make the return trip through the partner point.

Schematic of the connections between the top and bottom surface of a crystal
New research from Princeton University demonstrates the bizarre movement of electrons through a novel material called a Weyl semi-metal. The image shows a schematic of the connections at special values of electron momentum, which come in pairs and are called Weyl points (red and blue dots). An electron that leaves the surface on a red point can return through its partner blue point, and vice versa. This bizarre behavior is due to the “topological” connections through the bulk of the material. Image credit: Yazdani et al., Princeton University.

The team decided to explore the behavior of these electrons following research, published in Science last year by another Princeton team and separately by two independent groups, revealing that electrons in Weyl semi-metals are quite unusual. For example, their experiments implied that while most surface electrons create a wave pattern that resembles the spreading rings that ripple out when a stone is thrown into a pond, the surface electrons in the new materials should make only a half circle, earning them the name “Fermi arcs.”

To get a more direct look at the patterns of electron flow in Weyl semi-metals, postdoctoral researcher Hiroyuki Inoue and graduate student András Gyenis in Yazdani’s lab, with help from graduate student Seong Woo Oh, used a highly sensitive instrument called a scanning tunneling microscope, one of the few tools that can observe electron waves on a crystal surface.

They obtained the tantalum arsenide crystals from graduate student Shan Jiang and assistant professor Ni Ni at the University of California-Los Angeles.

The results were puzzling. “Some of the interference patterns that we expected to see were missing,” Yazdani said.

Comparison of experiment (left) to theory (right)
Comparison of experiment (left) to theory (right): The image on the left shows a pattern of waves imaged with scanning tunneling microscopy, revealing all the possible ways in which electrons can interfere with each other on the surface. This pattern is dictated by the connection of the surface electrons with the interior of the crystal, which is determined by the Weyl momentum of electrons, the special momentum when the electrons sink easily through the sample. The observed pattern closely matches the pattern predicted by theoretical calculations (right). Image credit: Yazdani et al., Princeton University.

To help explain the phenomenon, Yazdani consulted B. Andrei Bernevig, associate professor of physics at Princeton, who has expertise in the theory of topological materials and whose group was involved in the first predictions of Weyl semi-metals in a 2015 paper published in Physical Review X.

Bernevig, with help from postdoctoral researchers Jian Li and Zhijun Wang, realized that the observed pattern made sense if the electrons in these unusual materials were sinking into the bulk of the crystal. “Nobody had predicted that there would be signals of this type of transport from a scanning tunneling microscope, so it came as a bit of a surprise,” said Bernevig.

The next step, said Bernevig, is to look for the behavior in other crystals.

The research at Princeton was supported by the Army Research Office Multidisciplinary University Research Initiative program (W911NF-12-1-0461), the Gordon and Betty Moore Foundation as part of EPiQS initiative (GBMF4530), the National Science Foundation Materials Research Science and Engineering Centers programs through the Princeton Center for Complex Materials (DMR-1420541, NSF-DMR-1104612, NSF CAREER DMR-0952428), the David and Lucile Packard Foundation, and the W. M. Keck Foundation. This project was also made possible through use of the facilities at Princeton Nanoscale Microscopy Laboratory (ARO-W911NF-1-0262, ONR-N00014-14-1-0330, ONR-N00014-13-10661), the U.S. Department of Energy Basic Energy Sciences (DOE-BES) Defense Advanced Research Projects Agency, the U.S. Space and Naval Warfare Systems Command Meso program (N6601-11-1-4110, LPS and ARO-W911NF-1-0606), and the Eric and Wendy Schmidt Transformative Technology Fund at Princeton. Work at University of California–Los Angeles was supported by the DOE-BES (DE-SC0011978).

Read the abstract.

The article, “Quasiparticle interference of the Fermi arcs and surface-bulk connectivity of a Weyl semimetal,” by Hiroyuki Inoue, András Gyenis, Zhijun Wang, Jian Li, Seong Woo Oh, Shan Jiang, Ni Ni, B. Andrei Bernevig,and Ali Yazdani, was published in the March 11, 2016 issue of the journal Science.

Further reading:

M. Weng, C. Fang, Z. Fang, B. A. Bernevig, X. Dai, Phys. Rev. X 5, 011029 (2015)

Q. Lv et al., Phys. Rev. X 5, 1–8 (2015)

X. Yang et al., Nat. Phys. 11, 728–732 (2015)

Y. Xu et al., Science 349, 613–617 (2015)

How an artificial protein rescues dying cells (PNAS)

By Tien Nguyen, Department of Chemistry

A new study from Princeton has revealed how a synthetic protein revives E. coli cells that lack a life-sustaining gene, offering insight into how life can adapt to survive and potentially be reinvented.

Michael Hecht and Katie Digianantonio

Researchers in the Hecht lab discovered the unexpected way in which a synthetic protein called SynSerB promotes the growth of cells that lack the natural SerB gene, which encodes an enzyme responsible for the last step in the production of the essential amino acid serine. The findings were published in the Proceedings of the National Academy of Sciences.

The Hecht group first discovered SynSerB’s ability to rescue serine-depleted E. Coli cells in 2011. At that time, they also discovered several other de novo proteins capable of rescuing the deletions of three other essential proteins in E. coli. “These are novel proteins that have never existed on Earth, and aren’t related to anything on Earth yet they enable life to grow where it otherwise would not,” said Michael Hecht, professor of chemistry at Princeton and corresponding author on the article.

Natural proteins are complex molecular machines constructed from a pool of twenty different amino acids. Typically they range from several dozen to several hundred amino acids in length.  In principle, there are more possible protein sequences than atoms in the universe, but through evolution Nature has selected just a small fraction to carry out the cellular functions that make life possible.

“Those proteins must be really special,” Hecht said. “The driving question was, ‘Can we do that in the laboratory? Can we come up with non-natural sequences that are that special, from an enormous number of possibilities?”

To address this question, the Hecht lab developed a library of non-natural proteins guided by a concept called binary design. The idea was to narrow down the number of possible sequences by choosing from eleven select amino acids that were divided into two groups: polar and non-polar. By using only the polar or non-polar characteristics of those amino acids, the researchers could design a plethora of novel proteins to fold into a particular shape based on their affinity to and repulsion from water. Then, by allowing the specific positions to have different amino acids, the researchers were able to produce a diverse library of about one million proteins, each 102 amino acids long.

“We had to focus on certain subsets of proteins that we knew would fold and search there first for function,” said Katie Digianantonio, a graduate student in the Hecht lab and first author on the paper. “It’s like instead of searching the whole universe for life, we’re looking in specific solar systems.”

Having found several non-natural proteins that could rescue specific cell lines, this latest work details their investigation specifically into how SynSerB promotes cell growth. The most obvious explanation, that SynSerB simply catalyzed the same reaction performed by the deleted SerB gene, was discounted by an early experiment.

To discern SynSerB’s mechanism among the multitude of complex biochemical pathways in the cell, the researchers turned to a technique called RNA sequencing. This technique allowed them to take a detailed snapshot of the serine-depleted E. Coli cells with and without their synthetic protein and compare the differences.

“Instead of guessing and checking, we wanted to look at the overall environment to see what was happening,” Digianantonio said. The RNA sequencing experiment revealed that SynSerB induced overexpression of a protein called HisB, high levels of which have been shown to promote the key reaction normally performed by the missing gene. By enlisting the help of HisB, the non-natural protein was able to induce the production of serine, which ultimately allowed the cell to survive.

“Life is opportunistic. Some proteins are going to work by acting similarly to what they replaced and some will find another pathway,” Hecht said. “Either way it’s cool.”

Read the full articles here:

Digianantonio, K. M.; Hecht, M. H. “A protein constructed de novo enables cell growth by altering gene regulation.Proc. Natl. Acad. Sci. Published online before print on Feb. 16, 2016.

Fisher, M. A.; McKinley, K. L.; Bradley, L. H; Viola, S. R; Hecht, M. H. “De Novo designed proteins from a library of artificial sequences function in Escherichia Coli and enable cell growth.” PLoS One 2011, 6(1): e15364.

This research was funded by the US National Science Foundation (NSF) Grant MCB-1050510.