Controlling Temperature

Overview

Teaching: 20 min
Exercises: 0 min
Questions
  • What is temperature on the molecular level?

  • Why do we want to control the temperature of our MD-simulations?

  • What temperature control algorithms are commonly used?

  • What are the strengths and weaknesses of these common thermostats?

Objectives
  • Remind us how Maxwell-Boltzmann distributions relate to temperature.

  • Quickly review thermodynamic ensembles.

  • Learn about different thermostats and get an idea how they work.

  • Learn where thermostats that don’t produce correct thermodynamic ensembles can still be very useful.

Temperature at the molecular level.

Temperature is a macroscopic property and thermodynamics tells us that on the molecular level, the temperature of a system is defined by the average kinetic energy of all the particles (atoms, molecules) that make up the system. Statistical thermodynamics tells us that a system that contains a certain amount of energy and is in equilibrium, will have all of its energy distributed in the most probable way. This distribution is called the Maxwell-Boltzmann distribution. This means that in a system of particles, these particles won’t all have the same velocity but rather the velocities will follow a distribution that depends on their mass and the temperature of the system.

The speed distribution of ideal gas obeys the following relationship:

\[f_v(v)=\left(\frac{m}{2\pi k_B T}\right)^{3/2}\cdot4\pi v^2\cdot\exp({-mv^2/2k_B T})\]

Maxwell-Boltzmann distributions

Krypton at different temperatures   Noble gases with different mass at 298K

Plot of velocity distributions

The plot above shows how the Maxwell-Boltzmann distributions of ideal gas depend on temperature (left) and mass (right). The python code is in: code/thermostats/MB_dist_gas.py

Velocity distributions obtained from MD simulations of water at different temperature.

Plot of Maxwell-Boltzmann distributions

The plot above shows the distributions for the velocities of oxygen atoms from water molecules that have been simulated at three different temperatures (280K, 320K and 360K).

Thermodynamic ensembles

MD simulations are usually simulate one of the following thermodynamic ensembles:

  1. The microcanonical or constant NVE ensemble, where the number of particles (N), the system’s Volume (V) and the energy (E) are kept constant during the simulation.
  2. The canonical or constant NVT ensemble, where the number of particles, the Volume and the temperature (T) are kept constant.
  3. The isothermal-isobaric or NPT ensemble, where number of particles, the system’s pressure (P) and temperature are kept constant.
  1. The microcanonical or constant NVE ensemble.
  2. The canonical or constant NVT ensemble.
  3. The isothermal-isobaric or constant NPT ensemble.

Simulation of NVE ensemble is relatively easy to achieve, as long as the MD code manages to conserve the energy of the system. The NVT ensemble is more practically relevant, as in the real world it is much easier to manage the temperature of a system rather than it’s energy. To achieve this in our simulation, we need to use a temperature control algorithm, which is commonly called a thermostat. As many experiments in the lab are carried out at constant ambient pressure, rather than in a fixed/confined volume, the NPT ensemble is also widely used for simulations. In addition to using a thermostat to control the temperature, we also need to use a pressure control algorithm, often called a barostat.

The grand canonical ensemble however, where the chemical potential μ is kept constant, requires that the number of particles is allowed to change, which is not supported by most MD packages.

Temperature Control Algorithms

Despite their benefits, temperature control algorithms or thermostats have been described as “necessary evils” simply because they all introduce some artifacts. [Wong-ekkabut-2016].

To maintain a constant temperature, a thermostat allows energy to enter and leave the simulated system. Practically, thermostats do this by modifying velocities of subsets of particles. Temperature control methods can be divided into several categories:

  1. Strong coupling methods
  2. Weak coupling methods
  3. Stochastic methods
  4. Extended system dynamics

1. Strong coupling methods

Velocity rescaling

The idea is that we rescale the velocities at each step (or after a preset number of steps) so that we get the desired target temperature.

Velocity reassignment

Using this method, all of the velocities in the system are periodically reassigned (by assigning new randomized velocities) so that the entire system is set to the desired temperature.

Both methods do not generate the correct canonical ensemble. If the velocities are rescaled at every time step the kinetic energy will not fluctuate in time. This is inconsistent with statistical mechanics of the canonical ensemble. These methods are not recommended for equilibrium dynamics, but they are useful for system heating or cooling. Both have advantages and downsides. If you want to heat up a system, then rescaling will make hot spots even hotter, and this is not desired. Temperature reassignment avoids this problem, but the kinetic energy of particles is no longer consistent with their potential energy, and thus needs to be redistributed.

Downsides:

2. Weak coupling methods

Berendsen thermostat

The Berendsen thermostat at every simulation step rescales the velocities of all particles to remove a predefined fraction of the difference from the predefined temperature. [Berendsen-1984. Conceptually this thermostat works analogous to coupling the simulated system to a fictitious heat bath kept at some constant temperature. The rate of temperature equilibration with this thermostat is controlled by strength of the coupling. This makes the Berendsen thermostat a predictably converging and robust thermostat, which can be very useful when allowing the system to relax, for example when starting a molecular dynamics simulation after after an energy minimization.

However, a drawback of Berendsen thermostat is that it cannot be mapped onto a specific thermodynamic ensemble. It has been shown that it produces an energy distribution with a lower variance than of a true canonical ensemble because it disproportionally samples kinetic energies closer to \(T_0\) than would be observed in the true Maxwell-Boltzmann distribution [Basconi-2013 and Shirts-2013]. Therefore, the Berendsen thermostat should be avoided for production MD simulations in most cases.

Downsides:

Heat flows between the simulation system and the heat bath with the rate defined by a time constant \(\tau_T\) where small values of \(\tau_T\) mean tight coupling and fast temperature equilibration between the simulation and the bath. Large values of \(\tau_T\) mean slow equilibration and weak coupling. The time constant for heat bath coupling for the system is measured in picoseconds, and the default value is usually 1 ps.

Heat flows between the simulation system and the heat bath with the rate defined by a time constant \(\tau_T\)

3. Stochastic methods

Randomly assign a subset of atoms new velocities based on Maxwell-Boltzmann distributions for the target temperature. Randomization interferes with correlated motion and thus slows down the system’s kinetics.

Randomly assign a subset of atoms new velocities based on Maxwell-Boltzmann distributions for the target temperature. Randomization interferes with correlated motion and thus slows down the system’s kinetics.

Andersen thermostat

The Andersen thermostat controls the temperature by assigning a subset of atoms new velocities that are randomly selected from the Maxwell-Boltzmann distribution for the target temperature. The probability for a given particle to have it’s velocity reassigned at each step is small and can be expressed as \(\Delta t / \tau_T\) where \(\Delta t\) is the time step. This means that effectively on average every atom experiences a stochastic collision with a virtual particle every \(\Delta t\) [Andersen-1980]. A variant of the Andersen thermostat in which the velocities of all atoms are randomized every \(\Delta t\) is termed the “massive Andersen” thermostat [Basconi-2013].

The Andersen thermostat correctly samples the canonical ensemble, however momentum is not conserved with this thermostat. Due to velocity randomization it can impair some correlated motions and thus slow down the kinetics of the system. This algorithm therefore is not recommended when studying kinetics or diffusion properties of the system. This applies to all stochastic methods.

The number of steps between randomization of velocities to a distribution is an important parameter. Too high a collision rate (short interval between randomization) will slow down the speed at which the molecules explore configuration space, whereas too low a rate (long interval between randomization) means that the canonical distribution of energies will be sampled slowly.

The Lowe-Andersen thermostat

A variant of the Andersen thermostat that conserves momentum. This method perturbs the system dynamics to a far less than the original Andersen method. This alleviates suppressed diffusion in the system. [Koopman-2006].

Bussi stochastic velocity rescaling thermostat

Extension of the Berendsen method corrected for sampling the canonical distribution. The velocities of all the particles are rescaled by a properly chosen random factor. This thermostat produces a correct velocity distribution for the canonical ensemble, while retaining advantage of the Berendsen method that the temperature deviations from the target will fall via a first order exponential decay and is therefore avoiding oscillations that can be observed with extended system thermostats, like Nosé-Hoover. For most temperature-controlled MD-simulation, this thermostat is an excellent choice. [Bussi-2007]

Langevin thermostat

Langevin dynamics mimics the viscous aspect of a solvent and interaction with the environment by adding two force terms to the equation of motion: a frictional force and a random force. The frictional force and the random force combine to give the correct canonical ensemble.

The amount of friction is controlled by the damping coefficient. If its value is high, atoms will experience too much unnatural friction, however if the coefficient is too low, your system will fluctuate too far away from the desired temperature. The default value is usually 1/ps.

4. Extended system thermostats

Nosé-Hoover thermostat

The extended system method originally introduced by Nose and subsequently developed by Hoover. The idea is to consider the heat bath as an integral part of the system by addition of an artificial variable associated with a fictional “heat bath mass” to the equations of motion. An important feature of this method is that the temperature can be controlled without involving random numbers. Thus correlated motions are not impaired and this method describes kinetics and diffusion properties better. Because the time-evolution of the added variable is described by a second-order equation, heat may flow in and out of the system in an oscillatory fashion, leading to nearly periodic temperature fluctuations with the frequency proportional to the “heat bath mass”. Nose-1984, Hoover-1985.

The drawback of this thermostat is that it was shown to impart the canonical distribution as well as ergodicity (space-filling) of the thermostatted system.

Drawbacks:

The time constant parameter in this thermostat controls the period of temperature fluctuations at equilibrium.

Nosé-Hoover-chains

A modification of the Nosé-Hoover thermostat which includes not a single thermostat variable but a chain of variables. Martyna-1992. Nose-Hoover thermostat with one variable does not guarantee ergodicity, especially for small or stiff systems.(In an ergodic moving system any point, will eventually visit all parts of the space it moves in, in a uniform and random manner). Chaining variables behaves better for small or stiff cases, however an infinite chain is required to completely correct these issues.

Global and local thermostats

Global thermostats control temperature of all atom in a system uniformly. This may lead to cold solute and hot solvent due to a slow heat transfer.

With local thermostats it is possible to control temperature in selected groups of atoms independently. This works well for large solutes, but if solute is small this approach may result in large fluctuations and hence unphysical dynamics.

Specifying local thermostats

With NAMD it is possible to set coupling coefficients for each atom in occupancy or beta column of a pdb file:

langevinFile
langevinCol

tCoupleFile
tCoupleFCol

In GROMACS temperature of a selected groups of atoms can be controlled independently using tc-grps. Temperature coupling groups are coupled separately to temperature bath.

Selecting thermostats in molecular dynamics packages

Thermostat/MD package GROMACS NAMD AMBER
velocity rescaling   reascaleFreq (steps)  
velocity reassignment   reassignFreq (steps)  
Andersen tcoupl = andersen    
massive-Andersen tcoupl = andersen-massive   ntt = 2
Lowe-Andersen   loweAndersen on  
Berendsen tcoupl = berendsen tCouple on ntt = 1
Langevin   langevin on ntt = 3
Bussi tcoupl = V-rescale stochRescale on  
Nose-Hoover tcoupl = nose-hoover    
Nose-Hoover-chains nh-chain-length (default 10)    

Key Points

  • On the molecular level, temperature manifests itself as a number of particles having a certain average kinetic energy.

  • Thermodynamic ensembles are … FIXME

  • Some temperature control algorithms (e.g. the Berendsen thermostat) fail to produce kinetic energy distributions that represent a correct thermodynamic ensemble.

  • Other thermostats, like Nosé-Hoover, produce correct thermodynamic ensembles but can take long to converge.

  • Even though the the Berendsen thermostat fails to produce correct thermodynamic ensembles, it can be useful for system relaxation as it is robust and converges fast.