Provided by: python3-mpi4py-fft-doc_2.0.3-3build2_all bug

NAME

       mpi4py-fft - mpi4py-fft Documentation

MPI4PY-FFT

       Documentation Status

       mpi4py-fft is a Python package for computing Fast Fourier Transforms (FFTs).  Large arrays
       are distributed and communications are handled under the hood by MPI for Python  (mpi4py).
       To distribute large arrays we are using a new and completely generic algorithm that allows
       for any index set of a multidimensional array to be distributed. We  can  distribute  just
       one  index  (a slab decomposition), two index sets (pencil decomposition) or even more for
       higher-dimensional arrays.

       In mpi4py-fft there is also  included  a  Python  interface  to  the  FFTW  library.  This
       interface can be used without MPI, much like pyfftw, and even for real-to-real transforms,
       like discrete cosine or sine transforms.

   Introduction
       The Python package mpi4py-fft is a tool primarily for working with Fast Fourier Transforms
       (FFTs)  of  (large)  multidimensional arrays. There is really no limit as to how large the
       arrays can be, just as long as there is sufficient computing powers available. Also, there
       are  no  limits  as  to  how  transforms  can be configured. Just about any combination of
       transforms from the FFTW library is supported. Finally, mpi4py-fft can also be used simply
       to  distribute  and  redistribute  large  multidimensional  arrays  with  MPI, without any
       transforms at all.

       The main contribution of mpi4py-fft can be found  in  just  a  few  classes  in  the  main
       modules:

          • mpifftpencildistarraylibfftfftw

       The mpifft.PFFT class is the major entry point for most users. It is a highly configurable
       class, which under the  hood  distributes  large  dataarrays  and  performs  any  type  of
       transform, along any axes of a multidimensional array.

       The  pencil  module  is responsible for global redistributions through MPI.  However, this
       module is rarely used on its own, unless one simply needs  to  do  global  redistributions
       without any transforms at all. The pencil module is used heavily by the PFFT class.

       The  distarray  module  contains  classes for simply distributing multidimensional arrays,
       with no regards to transforms. The distributed arrays created from the  classes  here  can
       very  well  be  used  in  any  MPI  application  that  requires  a  large multidimensional
       distributed array.

       The libfft module provides a common interface to any of the serial transforms in the  FFTW
       library.

       The  fftw  module  contains  wrappers  to  the transforms provided by the FFTW library. We
       provide our own wrappers mainly because pyfftw does not include support  for  real-to-real
       transforms.  Through  the  interface  in  fftw  we  can  do  here,  in Python, pretty much
       everything that you can do in the original FFTW library.

   Global Redistributions
       In high performance computing large multidimensional  arrays  are  often  distributed  and
       shared amongst a large number of different processors.  Consider a large three-dimensional
       array of double (64 bit) precision and global shape (512, 1024, 2048). To lift this  array
       into  RAM  requires  8  GB of memory, which may be too large for a single, non-distributed
       machine. If, however, you have access to a distributed architecture,  you  can  split  the
       array  up and share it between, e.g., four CPUs (most supercomputers have either 2 or 4 GB
       of memory per CPU), which will only need to hold 2 GBs of the global array each. Moreover,
       many  algorithms  with  varying  degrees of locality can take advantage of the distributed
       nature of the array to compute local array  pieces  concurrently,  effectively  exploiting
       multiple processor resources.

       There  are  several  ways  of  distributing  a  large  multidimensional  array.  Two  such
       distributions for our three-dimensional global array (using 4 processors) are shown  below
       [image] [image]

       Here each color represents one of the processors. We note that in the first image only one
       of the three axes is distributed, whereas in the second  two  axes  are  distributed.  The
       first  configuration  corresponds  to  a  slab, whereas the second corresponds to a pencil
       distribution. With either distribution only one quarter of the large, global  array  needs
       to  be  kept  in  rapid  (RAM)  memory  for  each processor, which is great. However, some
       operations may then require data that is not available locally in its quarter of the total
       array. If that is so, the processors will need to communicate with each other and send the
       necessary data where it is needed. There are many such MPI routines designed  for  sending
       and receiving data.

       We  are  generally  interested in algorithms, like the FFT, that work on the global array,
       along one axis at the time. To be able to execute such algorithms, we need  to  make  sure
       that  the  local  arrays  have  access  to all of its data along this axis. For the figure
       above, the slab distribution gives each processor data that is fully available  along  two
       axes,  whereas  the  pencil  distribution  only  has  data fully available along one axis.
       Rearranging data, such that it becomes aligned in a different direction, is usually termed
       a  global  redistribution,  or  a global transpose operation. Note that with mpi4py-fft we
       always require that at  least  one  axis  of  a  multidimensional  array  remains  aligned
       (non-distributed).

       Distribution  and  global  redistribution is in mpi4py-fft handled by three classes in the
       pencil module:

          • PencilSubcommTransfer

       These classes are the low-level backbone of the higher-level PFFT and  DistArray  classes.
       To  use  these  low-level  classes  directly is not recommended and usually not necessary.
       However, for clarity we start by describing how these low-level classes work together.

       Lets first consider a 2D dataarray of global shape (8, 8) that will be  distributed  along
       axis 0. With a high level API we could then simply do:

          import numpy as np
          from mpi4py_fft import DistArray
          N = (8, 8)
          a = DistArray(N, [0, 1])

       where  the  [0, 1] list decides that the first axis can be distributed, whereas the second
       axis is using one processor only and as such is  aligned  (non-distributed).  We  may  now
       inspect the low-level Pencil class associated with a:

          p0 = a.pencil

       The  p0  Pencil  object  contains  information about the distribution of a 2D dataarray of
       global shape (8, 8). The distributed array a has been created using the  information  that
       is in p0, and p0 is used by a to look up information about the global array, for example:

          >>> a.alignment
          1
          >>> a.global_shape
          (8, 8)
          >>> a.subcomm
          (<mpi4py.MPI.Cartcomm at 0x10cc14a68>, <mpi4py.MPI.Cartcomm at 0x10e028690>)
          >>> a.commsizes
          [1, 1]

       Naturally,  the sizes of the communicators will depend on the number of processors used to
       run the program. If we used 4, then a.commsizes would return [1, 4].

       We note that a low-level approach to creating such a distributed array would be:

          import numpy as np
          from mpi4py_fft import Pencil, Subcomm
          from mpi4py import MPI
          comm = MPI.COMM_WORLD
          N = (8, 8)
          subcomm = Subcomm(comm, [0, 1])
          p0 = Pencil(subcomm, N, axis=1)
          a0 = np.zeros(p0.subshape)

       Note that this last array a0 would be equivalent to a, but it would be a pure Numpy  array
       (created  on  each  processor)  and  it would not contain any of the information about the
       global array that it is part of (global_shape, pencil, subcomm,  etc.).  It  contains  the
       same amount of data as a though and a0 is as such a perfectly fine distributed array. Used
       together with p0 it contains exactly the same information as a.

       Since at least one axis needs to be aligned (non-distributed), a  2D  array  can  only  be
       distributed  with  one processor group. If we wanted to distribute the second axis instead
       of the first, then we would have done:

          a = DistArray(N, [1, 0])

       With the low-level approach we would have had to use axis=0 in the creation of p0, as well
       as  [1,  0]  in  the  creation  of subcomm.  Another way to get the second pencil, that is
       aligned with axis 0, is to create it from p0:

          p1 = p0.pencil(0)

       Now the p1 object will represent a (8, 8) global array distributed in the second axis.

       Lets create a complete script (pencils.py) that fills the array a with the value  of  each
       processors  rank  (note  that  it would also work to follow the low-level approach and use
       a0):

          import numpy as np
          from mpi4py_fft import DistArray
          from mpi4py import MPI
          comm = MPI.COMM_WORLD
          N = (8, 8)
          a = DistArray(N, [0, 1])
          a[:] = comm.Get_rank()
          print(a.shape)

       We can run it with:

          mpirun -np 4 python pencils.py

       and obtain the printed results from the last line (print(a.shape)):

          (2, 8)
          (2, 8)
          (2, 8)
          (2, 8)

       The shape of the local a arrays is (2, 8) on all 4 processors. Now  assume  that  we  need
       these  data  aligned  in  the  x-direction (axis=0) instead. For this to happen we need to
       perform a global redistribution. The easiest approach is then to execute the following:

          b = a.redistribute(0)
          print(b.shape)

       which would print the following:

          (8, 2)
          (8, 2)
          (8, 2)
          (8, 2)

       Under the hood the global redistribution is executed with the help of the Transfer  class,
       that  is designed to transfer data between any two sets of pencils, like those represented
       by p0 and p1. With low-level API a transfer object may be created using  the  pencils  and
       the datatype of the array that is to be sent:

          transfer = p0.transfer(p1, np.float)

       Executing the global redistribution is then simply a matter of:

          a1 = np.zeros(p1.subshape)
          transfer.forward(a, a1)

       Now  it is important to realise that the global array does not change. The local a1 arrays
       will now contain the same data as a, only aligned differently.  However, the  exchange  is
       not  performed  in-place.  The new array is as such a copy of the original that is aligned
       differently.  Some images, Fig. %s and Fig. %s, can be used to illustrate:
         [image] Original 4  pencils  (p0)  of  shape  (2,  8)  aligned  in   y-direction.  Color
         represents rank..UNINDENT
         [image]  4 pencils (p1) of shape (8, 2) aligned in x-direction after receiving data from
         p0. Data is the same as in Fig. %s, only aligned differently..UNINDENT

         Mathematically, we will denote the entries of a two-dimensional global array as  u_{j_0,
         j_1},  where  j_0\in  \textbf{j}_0=[0,  1, \ldots, N_0-1] and j_1\in \textbf{j}_1=[0, 1,
         \ldots, N_1-1]. The shape of the array is then (N_0, N_1). A global array  u_{j_0,  j_1}
         distributed in the first axis (as shown in Fig. %s) by processor group P, containing |P|
         processors, is denoted as

                                              u_{j_0/P, j_1}

         The global redistribution, from alignment in axis 1 to alignment in axis 0, as from Fig.
         %s to Fig. %s above, is denoted as

                       u_{j_0, j_1/P} \xleftarrow[P]{1\rightarrow 0} u_{j_0/P, j_1}

         This operation corresponds exactly to the forward transfer defined above:

          transfer.forward(a0, a1)

       If we need to go the other way

                      u_{j_0/P, j_1} \xleftarrow[P]{0\rightarrow 1} u_{j_0, j_1/P}

       this corresponds to:

          transfer.backward(a1, a0)

       Note  that  the  directions  (forward/backward) here depends on how the transfer object is
       created. Under the hood all transfers are executing calls to MPI.Alltoallw.

   Multidimensional distributed arrays
       The  procedure  discussed  above  remains  the  same  for  any  type  of  array,  of   any
       dimensionality.  With  mpi4py-fft  we can distribute any array of arbitrary dimensionality
       using any number of processor groups. We only require that the number of processor  groups
       is  at  least  one less than the number of dimensions, since one axis must remain aligned.
       Apart from this the distribution is completely configurable through  the  classes  in  the
       pencil module.

       We   denote  a  global  d-dimensional  array  as  u_{j_0,  j_1,  \ldots,  j_{d-1}},  where
       j_m\in\textbf{j}_m for m=[0, 1, \ldots, d-1].  A d-dimensional array distributed with only
       one  processor  group  in the first axis is denoted as u_{j_0/P, j_1, \ldots, j_{d-1}}. If
       using more than one processor group, the groups are indexed, like P_0, P_1 etc.

       Lets illustrate using a 4-dimensional array with 3 processor  groups.  Let  the  array  be
       aligned  only  in axis 3 first (u_{j_0/P_0, j_1/P_1, j_2/P_2, j_3}), and then redistribute
       for alignment along axes 2, 1 and finally 0. Mathematically, we will now be executing  the
       three following global redistributions:

       u_{j_0/P_0, j_1/P_1, j_2, j_3/P_2} \xleftarrow[P_2]{3 \rightarrow 2}  u_{j_0/P_0, j_1/P_1,
                                            j_2/P_2, j_3} \\
       u_{j_0/P_0, j_1, j_2/P_1, j_3/P_2} \xleftarrow[P_1]{2 \rightarrow 1}  u_{j_0/P_0, j_1/P_1,
       j_2,  j_3/P_2}  \\  u_{j_0,  j_1/P_0,  j_2/P_1, j_3/P_2} \xleftarrow[P_0]{1 \rightarrow 0}
       u_{j_0/P_0, j_1, j_2/P_1, j_3/P_2}

       Note that in the first step it  is  only  processor  group  P_2  that  is  active  in  the
       redistribution, and the output (left hand side) is now aligned in axis 2. This can be seen
       since there is no processor group there to share  the  j_2  index.   In  the  second  step
       processor group P_1 is the active one, and in the final step P_0.

       Now,  it  is  not  necessary  to  use  three  processor  groups  just  because  we  have a
       four-dimensional array. We could just as well have been using 2 or  1.  The  advantage  of
       using  more  groups  is that you can then use more processors in total. Assuming N = N_0 =
       N_1 = N_2 = N_3, you can use a maximum of  N^p  processors,  where  p  is  the  number  of
       processor  groups.  So for an array of shape (8,8,8,8) it is possible to use 8, 64 and 512
       number of processors for 1, 2 and 3 processor groups, respectively. On the other hand,  if
       you  can  get  away with it, or if you do not have access to a great number of processors,
       then fewer groups are usually found to be faster for the  same  number  of  processors  in
       total.

       We can implement the global redistribution using the high-level DistArray class:

          N = (8, 8, 8, 8)
          a3 = DistArray(N, [0, 0, 0, 1])
          a2 = a3.redistribute(2)
          a1 = a2.redistribute(1)
          a0 = a1.redistribute(0)

       Note that the three redistribution steps correspond exactly to the three steps in (1).

       Using  a  low-level  API  the same can be achieved with a little more elaborate coding. We
       start by creating pencils for the 4 different alignments:

          subcomm = Subcomm(comm, [0, 0, 0, 1])
          p3 = Pencil(subcomm, N, axis=3)
          p2 = p3.pencil(2)
          p1 = p2.pencil(1)
          p0 = p1.pencil(0)

       Here we have defined 4 different pencil groups, p0, p1, p2, p3, aligned in axis  0,  1,  2
       and 3, respectively. Transfer objects for arrays of type np.float are then created as:

          transfer32 = p3.transfer(p2, np.float)
          transfer21 = p2.transfer(p1, np.float)
          transfer10 = p1.transfer(p0, np.float)

       Note  that  we  can create transfer objects between any two pencils, not just neighbouring
       axes. We may now perform three different global redistributions as:

          a0 = np.zeros(p0.subshape)
          a1 = np.zeros(p1.subshape)
          a2 = np.zeros(p2.subshape)
          a3 = np.zeros(p3.subshape)
          a0[:] = np.random.random(a0.shape)
          transfer32.forward(a3, a2)
          transfer21.forward(a2, a1)
          transfer10.forward(a1, a0)

       Storing this code under pencils4d.py, we  can  use  8  processors  that  will  give  us  3
       processor groups with 2 processors in each group:

          mpirun -np 8 python pencils4d.py

       Note that with the low-level approach we can now easily go back using the reverse backward
       method of the Transfer objects:

          transfer10.backward(a0, a1)

       A different approach is also possible with the high-level API:

          a0.redistribute(out=a1)
          a1.redistribute(out=a2)
          a2.redistribute(out=a3)

       which corresponds to the backward transfers. However, with the high-level API the transfer
       objects are created (and deleted on exit) during the call to redistribute and as such this
       latter approach may be slightly less efficient.

   Discrete Fourier Transforms
       Consider  first  two  one-dimensional  arrays  \boldsymbol{u}  =  \{u_j\}_{j=0}^{N-1}  and
       \boldsymbol{\hat{u}}  =\{\hat{u}_k\}_{k=0}^{N-1}.  We  define  the  forward  and  backward
       Discrete Fourier transforms (DFT), respectively, as

       \hat{u}_k &= \frac{1}{N}\sum_{j=0}^{N-1}u_j e^{-2\pi i j k / N},  \quad  \forall  \,  k\in
                                    \textbf{k}=0, 1, \ldots, N-1, \\
       u_j &= \sum_{k=0}^{N-1}\hat{u}_k e^{2\pi i j k / N}, \quad \forall \, j\in\textbf{j}=0, 1,
       \ldots, N-1,

       where i=\sqrt{-1}. Discrete Fourier transforms are computed efficiently  using  algorithms
       termed Fast Fourier Transforms, known in short as FFTs.

       NOTE:
          The  index set for wavenumbers \textbf{k} is usually not chosen as [0, 1, \ldots, N-1],
          but \textbf{k}=[-N/2, -N/2-1, \ldots,  N/2-1]  for  even  N  and  \textbf{k}=[-(N-1)/2,
          -(N-1)/2+1,  \ldots,  (N-1)/2]  for odd N. See numpy.fft.fftfreq.  Also note that it is
          possible to tweak the default normalization used above when calling either  forward  or
          backward transforms.

       A  more  compact notation is commonly used for the DFTs, where the 1D forward and backward
       transforms are written as

                        \boldsymbol{\hat{u}} &= \mathcal{F}(\boldsymbol{u}), \\
       \boldsymbol{u} &= \mathcal{F}^{-1}(\boldsymbol{\hat{u}}).

       Numpy, Scipy, and many  other  scientific  softwares  contain  implementations  that  make
       working  with  Fourier series simple and straight forward. These 1D Fourier transforms can
       be implemented easily with just Numpy as, e.g.:

          import numpy as np
          N = 16
          u = np.random.random(N)
          u_hat = np.fft.fft(u)
          uc = np.fft.ifft(u_hat)
          assert np.allclose(u, uc)

       However, there is a minor difference. Numpy performs by default the 1/N scaling  with  the
       backward  transform (ifft) and not the forward as shown in (2). These are merely different
       conventions and not important as long as one is aware of them. We use the scaling  on  the
       forward  transform simply because this follows naturally when using the harmonic functions
       e^{i k x} as basis functions when solving PDEs with the spectral Galerkin  method  or  the
       spectral collocation method (see chap. 3).

       With  mpi4py-fft  the  same  operations  take  just  a  few more steps, because instead of
       executing ffts directly, like in the calls for np.fft.fft  and  np.fft.ifft,  we  need  to
       create the objects that are to do the transforms first. We need to plan the transforms:

          from mpi4py_fft import fftw
          u = fftw.aligned(N, dtype=np.complex)
          u_hat = fftw.aligned_like(u)
          fft = fftw.fftn(u, flags=(fftw.FFTW_MEASURE,))        # plan fft
          ifft = fftw.ifftn(u_hat, flags=(fftw.FFTW_ESTIMATE,)) # plan ifft
          u[:] = np.random.random(N)
          # Now execute the transforms
          u_hat = fft(u, u_hat, normalize=True)
          uc = ifft(u_hat)
          assert np.allclose(uc, u)

       The  planning  of transforms makes an effort to find the fastest possible transform of the
       given kind. See more in The fftw module.

   Multidimensional transforms
       It is for multidimensional arrays that it starts to become  interesting  for  the  current
       software. Multidimensional arrays are a bit tedious with notation, though, especially when
       the number of dimensions grow. We will stick with the index notation because  it  is  most
       stright forward in comparison with implementation.

       We  denote  the entries of a two-dimensional array as u_{j_0, j_1}, which corresponds to a
       row-major matrix  \boldsymbol{u}=\{u_{j_0,  j_1}\}_{(j_0,  j_1)  \in  \textbf{j}_0  \times
       \textbf{j}_1}   of   size   N_0\cdot   N_1.  Denoting  also  \omega_m=j_m  k_m  /  N_m,  a
       two-dimensional forward and backward DFT can be defined as

       \hat{u}_{k_0,k_1} &= \frac{1}{N_0}\sum_{j_0 \in \textbf{j}_0}\Big(  e^{-2\pi  i  \omega_0}
       \frac{1}{N_1}  \sum_{j_1\in  \textbf{j}_1}  \Big(  e^{-2\pi  i  \omega_1} u_{j_0,j_1}\Big)
              \Big), \quad \forall \, (k_0, k_1) \in \textbf{k}_0  \times \textbf{k}_1, \\
       u_{j_0,   j_1}   &=    \sum_{k_1\in    \textbf{k}_1}    \Big(    e^{2\pi    i    \omega_1}
       \sum_{k_0\in\textbf{k}_0} \Big(  e^{2\pi i \omega_0} \hat{u}_{k_0, k_1} \Big) \Big), \quad
       \forall \, (j_0, j_1) \in \textbf{j}_0 \times \textbf{j}_1.

       Note that the forward transform corresponds to taking the 1D Fourier transform first along
       axis  1,  once  for  each  of  the  indices  in \textbf{j}_0.  Afterwords the transform is
       executed along axis 0. The two steps are more easily understood if we break  things  up  a
       little bit and write the forward transform in (3) in two steps as

       \tilde{u}_{j_0,k_1}  &=  \frac{1}{N_1}\sum_{j_1  \in  \textbf{j}_1} u_{j_0,j_1} e^{-2\pi i
                          \omega_1}, \quad \forall \, k_1 \in \textbf{k}_1, \\
       \hat{u}_{k_0,k_1} &= \frac{1}{N_0}\sum_{j_0 \in \textbf{j}_0} \tilde{u}_{j_0,k_1} e^{-2\pi
       i \omega_0}, \quad \forall \, k_0 \in \textbf{k}_0.

       The backward (inverse) transform if performed in the opposite order, axis 0 first and then
       1. The order is actually  arbitrary,  but  this  is  how  is  is  usually  computed.  With
       mpi4py-fft the order of the directional transforms can easily be configured.

       We can write the complete transform on compact notation as

                        \boldsymbol{\hat{u}} &= \mathcal{F}(\boldsymbol{u}), \\
       \boldsymbol{u} &= \mathcal{F}^{-1}(\boldsymbol{\hat{u}}).

       But  if  we  denote  the  two  partial  transforms  along  each  axis as \mathcal{F}_0 and
       \mathcal{F}_1, we can also write it as

                \boldsymbol{\hat{u}} &= \mathcal{F}_0(\mathcal{F}_1(\boldsymbol{u})), \\
       \boldsymbol{u} &= \mathcal{F}_1^{-1}(\mathcal{F}_0^{-1}(\boldsymbol{\hat{u}})).

       Extension to multiple dimensions is straight forward. We denote a d-dimensional  array  as
       u_{j_0, j_1, \ldots, j_{d-1}} and a partial transform of u along axis i is denoted as

       \tilde{u}_{j_0, \ldots, k_i, \ldots, j_{d-1}} = \mathcal{F}_i(u_{j_0, \ldots, j_i, \ldots, j_d})

       We  get  the  complete  multidimensional  transforms  on short form still as (5), and with
       partial transforms as

       \boldsymbol{\hat{u}}           &=           \mathcal{F}_0(\mathcal{F}_1(            \ldots
                                 \mathcal{F}_{d-1}(\boldsymbol{u})), \\
       \boldsymbol{u}      &=      \mathcal{F}_{d-1}^{-1}(     \mathcal{F}_{d-2}^{-1}(     \ldots
       \mathcal{F}_0^{-1}(\boldsymbol{\hat{u}}))).

       Multidimensional transforms are straightforward to implement in Numpy

          import numpy as np
          M, N = 16, 16
          u = np.random.random((M, N))
          u_hat = np.fft.rfftn(u)
          uc = np.fft.irfftn(u_hat)
          assert np.allclose(u, uc)

   The fftw module
       The fftw module provides an interface to most of  the  FFTW  library.  In  the  fftw.xfftn
       submodule there are planner functions for:

          • fftn() - complex-to-complex forward Fast Fourier Transforms

          • ifftn() - complex-to-complex backward Fast Fourier Transforms

          • rfftn() - real-to-complex forward FFT

          • irfftn() - complex-to-real backward FFT

          • dctn() - real-to-real Discrete Cosine Transform (DCT)

          • idctn() - real-to-real inverse DCT

          • dstn() - real-to-real Discrete Sine Transform (DST)

          • idstn() - real-to-real inverse DST

          • hfftn() - complex-to-real forward FFT with Hermitian symmetry

          • ihfftn() - real-to-complex backward FFT with Hermitian symmetry

       All  these  transform  functions  return  instances of one of the classes fftwf_xfftn.FFT,
       fftw_xfftn.FFT or fftwl_xfftn.FFT, depending on  the  requested  precision  being  single,
       double  or  long  double,  respectively.  Except  from  precision,  the  tree  classes are
       identical.  All transforms are non-normalized by default. Note that  all  these  functions
       are  planners.  They  do  not  execute the transforms, they simply return an instance of a
       class that can do it (see docstrings of each function for usage).   For  quick  reference,
       the 2D transform shown for Numpy can be done using fftw as:

          from mpi4py_fft.fftw import rfftn as plan_rfftn, irfftn as plan_irfftn
          from mpi4py_fft.fftw import FFTW_ESTIMATE
          rfftn = plan_rfftn(u.copy(), flags=(FFTW_ESTIMATE,))
          irfftn = plan_irfftn(u_hat.copy(), flags=(FFTW_ESTIMATE,))
          u_hat = rfftn(uc, normalize=True)
          uu = irfftn(u_hat)
          assert np.allclose(uu, uc)

       Note  that  since  all  the  functions  in  the  above list are planners, an extra step is
       required in comparison with Numpy. Also note that we are using copies of the u  and  u_hat
       arrays  in creating the plans. This is done because the provided arrays will be used under
       the hood as work arrays for the rfftn() and irfftn() functions, and the work arrays may be
       destroyed upon creation.

       The  real-to-real  transforms  are by FFTW defined as one of (see definitions and extended
       definitions)

          • FFTW_REDFT00

          • FFTW_REDFT01

          • FFTW_REDFT10

          • FFTW_REDFT11

          • FFTW_RODFT00

          • FFTW_RODFT01

          • FFTW_RODFT10

          • FFTW_RODFT11

       Different real-to-real cosine and sine transforms may be combined into  one  object  using
       factory.get_planned_FFT()  with  a  list  of different transform kinds. However, it is not
       possible to combine, in one single object, real-to-real transforms  with  real-to-complex.
       For such transforms more than one object is required.

   Parallel Fast Fourier Transforms
       Parallel  FFTs  are  computed  through  a combination of global redistributions and serial
       transforms. In mpi4py-fft the interface to performing  such  parallel  transforms  is  the
       mpifft.PFFT  class.  The  class  is  highly  configurable and best explained through a few
       examples.

   Slab decomposition
       With slab decompositions we use only one group of processors and distribute only one index
       of a multidimensional array at the time.

       Consider  the  complete  transform  of a three-dimensional array of random numbers, and of
       shape (128, 128, 128). We can plan the transform of such an array with the following  code
       snippet:

          import numpy as np
          from mpi4py import MPI
          from mpi4py_fft import PFFT, newDistArray
          N = np.array([128, 128, 128], dtype=int)
          fft = PFFT(MPI.COMM_WORLD, N, axes=(0, 1, 2), dtype=np.float, grid=(-1,))

       Here the signature N, axes=(0, 1, 2), dtype=np.float, grid=(-1,) tells us that the created
       fft instance is planned such as to slab distribute (along first axis) and transform any 3D
       array  of  shape  N and type np.float. Furthermore, we plan to transform axis 2 first, and
       then 1 and 0, which is exactly the reverse order of axes=(0, 1,  2).  Mathematically,  the
       planned transform corresponds to

         \tilde{u}_{j_0/P,k_1,k_2} &= \mathcal{F}_1( \mathcal{F}_{2}(u_{j_0/P, j_1, j_2})), \\
       \tilde{u}_{j_0,  k_1/P,  k_2} &\xleftarrow[P]{1\rightarrow 0} \tilde{u}_{j_0/P, k_1, k_2},
       \\ \hat{u}_{k_0,k_1/P,k_2} &= \mathcal{F}_0(\tilde{u}_{j_0, k_1/P, k_2}).

       Note that axis 0 is distributed on the input array and axis 1 on the output array. In  the
       first  step  above  we  compute  the transforms along axes 2 and 1 (in that order), but we
       cannot compute the serial transform along axis 0 since the global array is distributed  in
       that direction. We need to perform a global redistribution, the middle step, that realigns
       the global data such that it is aligned in axes 0.  With data aligned in axis  0,  we  can
       perform the final transform \mathcal{F}_{0} and be done with it.

       Assume  now  that  all the code in this section is stored to a file named pfft_example.py,
       and add to the above code:

          u = newDistArray(fft, False)
          u[:] = np.random.random(u.shape).astype(u.dtype)
          u_hat = fft.forward(u, normalize=True) # Note that normalize=True is default and can be omitted
          uj = np.zeros_like(u)
          uj = fft.backward(u_hat, uj)
          assert np.allclose(uj, u)
          print(MPI.COMM_WORLD.Get_rank(), u.shape)

       Running this code with two processors (mpirun -np 2 python pfft_example.py)  should  raise
       no exception, and the output should be:

          1 (64, 128, 128)
          0 (64, 128, 128)

       This  shows  that  the first index has been shared between the two processors equally. The
       array u thus corresponds to  u_{j_0/P,j_1,j_2}.  Note  that  the  newDistArray()  function
       returns   a   DistArray  object,  which  in  turn  is  a  subclassed  Numpy  ndarray.  The
       newDistArray() function uses fft to determine the size and type of the created distributed
       array,  i.e.,  (64,  128,  128)  and  np.float  for  both  processors.  The False argument
       indicates that the shape and type should be that of the input array,  as  opposed  to  the
       output array type (\hat{u}_{k_0,k_1/P,k_2} that one gets with True).

       Note  that because the input array is of real type, and not complex, the output array will
       be of global shape:

          128, 128, 65

       The output array will be distributed in axis 1, so the output array shape should be  (128,
       64, 65) on each processor. We check this by adding the following code and rerunning:

          u_hat = newDistArray(fft, True)
          print(MPI.COMM_WORLD.Get_rank(), u_hat.shape)

       leading to an additional print of:

          1 (128, 64, 65)
          0 (128, 64, 65)

       To  distribute  in  the  first  axis  first  is default and most efficient for row-major C
       arrays. However, we can easily configure the fft instance by modifying the  axes  keyword.
       Changing for example to:

          fft = PFFT(MPI.COMM_WORLD, N, axes=(2, 0, 1), dtype=np.float)

       and  axis  1 will be transformed first, such that the global output array will be of shape
       (128, 65, 128). The distributed input and output arrays will now have shape:

          0 (128, 128, 64)
          1 (128, 128, 64)

          0 (64, 65, 128)
          1 (64, 65, 128)

       Note that the input array will be distributed in axis 2 and the output in axis 0.

       Another way to tweak the distribution is to use the Subcomm class directly:

          from mpi4py_fft.pencil import Subcomm
          subcomms = Subcomm(MPI.COMM_WORLD, [1, 0, 1])
          fft = PFFT(subcomms, N, axes=(0, 1, 2), dtype=np.float)

       Here the subcomms tuple will decide that axis 1 should be distributed,  because  the  only
       zero  in  the  list [1, 0, 1] is along axis 1. The ones determine that axes 0 and 2 should
       use one processor each, i.e., they should be non-distributed.

       The PFFT class has a few additional keyword arguments that one should  be  aware  of.  The
       default behaviour of PFFT is to use one transform object for each axis, and then use these
       sequentially.  Setting collapse=True will attempt to  minimize  the  number  of  transform
       objects  by combining whenever possible. Take our example, the array u_{j_0/P,j_1,j_2} can
       transform along both  axes  1  and  2  simultaneously,  without  any  intermediate  global
       redistributions. By setting collapse=True only one object of rfftn(u, axes=(1, 2)) will be
       used instead of two (like fftn(rfftn(u, axes=2), axes=1)).  Note that a collapse can  also
       be configured through the axes keyword, using:

          fft = PFFT(MPI.COMM_WORLD, N, axes=((0,), (1, 2)), dtype=np.float)

       will collapse axes 1 and 2, just like one would obtain with collapse=True.

       If  serial  transforms  other  than fftn()/rfftn() and ifftn()/irfftn() are required, then
       this can be achieved using the transforms keyword and a dictionary pointing from  axes  to
       the  type  of  transform.  We  can  for  example combine real-to-real with real-to-complex
       transforms like this:

          from mpi4py_fft.fftw import rfftn, irfftn, dctn, idctn
          import functools
          dct = functools.partial(dctn, type=3)
          idct = functools.partial(idctn, type=3)
          transforms = {(0,): (rfftn, irfftn), (1, 2): (dct, idct)}
          r2c = PFFT(MPI.COMM_WORLD, N, axes=((0,), (1, 2)), transforms=transforms)
          u = newDistArray(r2c, False)
          u[:] = np.random.random(u.shape).astype(u.dtype)
          u_hat = r2c.forward(u)
          uj = np.zeros_like(u)
          uj = r2c.backward(u_hat, uj)
          assert np.allclose(uj, u)

       As a more complex example consider a 5-dimensional array where for some reason you need to
       perform discrete cosine transforms in axes 1 and 2, discrete sine transforms in axes 3 and
       4, and a regular Fourier transform in the first axis.  Here it makes sense to collapse the
       (1,  2)  and  (3, 4) axes, which leaves only the first axis uncollapsed. Hence we can then
       only use one processor group and a slab decomposition, whereas without collapsing we could
       have used four groups.  A parallel transform object can be created and tested as:

          N = (5, 6, 7, 8, 9)
          dctn = functools.partial(fftw.dctn, type=3)
          idctn = functools.partial(fftw.idctn, type=3)
          dstn = functools.partial(fftw.dstn, type=3)
          idstn = functools.partial(fftw.idstn, type=3)
          fft = PFFT(MPI.COMM_WORLD, N, ((0,), (1, 2), (3, 4)), grid=(-1,),
                     transforms={(1, 2): (dctn, idctn), (3, 4): (dstn, idstn)})

          A = newDistArray(fft, False)
          A[:] = np.random.random(A.shape)
          C = fftw.aligned_like(A)
          B = fft.forward(A)
          C = fft.backward(B, C)
          assert np.allclose(A, C)

   Pencil decomposition
       A  pencil  decomposition uses two groups of processors. Each group then is responsible for
       distributing one index set each of a multidimensional array.   We  can  perform  a  pencil
       decomposition  simply by running the first example from the previous section, but now with
       4 processors. To remind you, we put this in pfft_example.py, where now grid=(-1,) has been
       removed in the PFFT calling:

          import numpy as np
          from mpi4py import MPI
          from mpi4py_fft import PFFT, newDistArray

          N = np.array([128, 128, 128], dtype=int)
          fft = PFFT(MPI.COMM_WORLD, N, axes=(0, 1, 2), dtype=np.float)
          u = newDistArray(fft, False)
          u[:] = np.random.random(u.shape).astype(u.dtype)
          u_hat = fft.forward(u)
          uj = np.zeros_like(u)
          uj = fft.backward(u_hat, uj)
          assert np.allclose(uj, u)
          print(MPI.COMM_WORLD.Get_rank(), u.shape)

       The output of running mpirun -np 4 python pfft_example.py will then be:

          0 (64, 64, 128)
          2 (64, 64, 128)
          3 (64, 64, 128)
          1 (64, 64, 128)

       Note that now both the two first index sets are shared, so we have a pencil decomposition.
       The shared input array is now denoted as u_{j_0/P_0,j_1/P_1,j2} and the  complete  forward
       transform performs the following 5 steps:

           \tilde{u}_{j_0/P_0,j_1/P_1,k_2} &= \mathcal{F}_{2}(u_{j_0/P_0, j_1/P_1, j_2}), \\
       \tilde{u}_{j_0/P_0,  j_1,  k_2/P_1}  &\xleftarrow[P_1]{2\rightarrow 1} \tilde{u}_{j_0/P_0,
       j_1/P_1, k_2},  \\  \tilde{u}_{j_0/P_0,k_1,k_2/P_1}  &=  \mathcal{F}_1(\tilde{u}_{j_0/P_0,
       j_1,  k_2/P_1}),  \\  \tilde{u}_{j_0,  k_1/P_0, k_2/P_1} &\xleftarrow[P_0]{1\rightarrow 0}
       \tilde{u}_{j_0/P_0,     k_1,     k_2/P_1},     \\     \hat{u}_{k_0,k_1/P_0,k_2/P_1}     &=
       \mathcal{F}_0(\tilde{u}_{j_0, k_1/P_0, k_2/P_1}).

       Like  for the slab decomposition, the order of the different steps is configurable. Simply
       change the value of axes, e.g., as:

          fft = PFFT(MPI.COMM_WORLD, N, axes=(2, 0, 1), dtype=np.float)

       and the input and output arrays will be of shape:

          3 (64, 128, 64)
          2 (64, 128, 64)
          1 (64, 128, 64)
          0 (64, 128, 64)

          3 (64, 32, 128)
          2 (64, 32, 128)
          1 (64, 33, 128)
          0 (64, 33, 128)

       We see that the input  array  is  aligned  in  axis  1,  because  this  is  the  direction
       transformed first.

   Convolution
       Working with Fourier one sometimes need to transform the product of two or more functions,
       like

       \widehat{ab}_k = \int_{0}^{2\pi} a b e^{-i k x} dx, \quad \forall k \in [-N/2, \ldots, N/2-1]

       computed with DFT as

       \widehat{ab}_k = \frac{1}{N}\sum_{j=0}^{N-1}a_j b_j e^{-2\pi i j k / N}, \quad \forall  \,
                                      k\in [-N/2, \ldots, N/2-1].

       NOTE:
          We are here assuming an even number N and use wavenumbers centered around zero.

       If a and b are two Fourier series with their own coefficients:

                           a &= \sum_{p=-N/2}^{N/2-1} \hat{a}_p e^{i p x}, \\
       b &= \sum_{q=-N/2}^{N/2-1} \hat{b}_q e^{i q x},

       then we can insert for the two sums from (11) in (9) and get

       \widehat{ab}_k  &=  \int_{0}^{2\pi}  \left(  \sum_{p=-N/2}^{N/2-1}  \hat{a}_p  e^{i  p  x}
       \sum_{q=-N/2}^{N/2-1} \hat{b}_q e^{i q x} \right)  e^{-i k x} dx, \quad \forall \,  k  \in
                                        [-N/2, \ldots, N/2-1] \\
       \widehat{ab}_k   &=   \sum_{p=-N/2}^{N/2-1}   \sum_{q=-N/2}^{N/2-1}  \hat{a}_p   \hat{b}_q
       \int_{0}^{2\pi} e^{-i (p+q-k) x} dx, \quad \forall \, k \in [-N/2, \ldots, N/2-1]

       The final integral is 2\pi for p+q=k and zero otherwise. Consequently, we get

       \widehat{ab}_k = 2\pi  \sum_{p=-N/2}^{N/2-1}\sum_{q=-N/2}^{N/2-1}  \hat{a}_p   \hat{b}_{q}
                     \delta_{p+q, k} , \quad \forall \, k \in [-N/2, \ldots, N/2-1]

       Unfortunately,  the  convolution  sum  (13)  is  very expensive to compute, and the direct
       application of (10) leads to aliasing errors.  Luckily  there  is  a  fast  approach  that
       eliminates aliasing as well.

       The  fast,  alias-free, approach makes use of the FFT and zero-padded coefficient vectors.
       The idea is to zero-pad \hat{a} and \hat{b}  in  spectral  space  such  that  we  get  the
       extended sums

                   A_j &= \sum_{p=-M/2}^{M/2-1} \hat{\hat{a}}_p e^{2 \pi i p j/M}, \\
       B_j &= \sum_{q=-M/2}^{M/2-1} \hat{\hat{b}}_q e^{2 \pi i q j/M},

       where M>N and where the coefficients have been zero-padded such that

                   \hat{\hat{a}}_p = \begin{cases} \hat{a}_p, &\forall |p| \le N/2 \\
                                       0, &\forall |p| \gt N/2 \end{cases}

       Now compute the nonlinear term in the larger physical space and compute the convolution as

       \widehat{ab}_k  =  \frac{1}{M} \sum_{j=0}^{M-1} A_j B_j e^{- 2 \pi i k j/M}, \quad \forall
                                     \, k \in [-M/2, \ldots, M/2-1]

       Finally, truncate the vector \widehat{ab}_k  to  the  original  range  k\in[-N/2,  \ldots,
       N/2-1], simply by eliminating all the wavenumbers higher than |N/2|.

       With  mpi4py-fft  we  can  compute  this convolution using the padding keyword of the PFFT
       class:

          import numpy as np
          from mpi4py_fft import PFFT, newDistArray
          from mpi4py import MPI

          comm = MPI.COMM_WORLD
          N = (128, 128)   # Global shape in physical space
          fft = PFFT(comm, N, padding=[1.5, 1.5], dtype=np.complex)

          # Create arrays in normal spectral space
          a_hat = newDistArray(fft, True)
          b_hat = newDistArray(fft, True)
          a_hat[:] = np.random.random(a_hat.shape) + np.random.random(a_hat.shape)*1j
          b_hat[:] = np.random.random(a_hat.shape) + np.random.random(a_hat.shape)*1j

          # Transform to real space with padding
          a = newDistArray(fft, False)
          b = newDistArray(fft, False)
          assert a.shape == (192//comm.Get_size(), 192)
          a = fft.backward(a_hat, a)
          b = fft.backward(b_hat, b)

          # Do forward transform with truncation
          ab_hat = fft.forward(a*b)

       NOTE:
          The padded instance of the PFFT class is often used in addition to a regular non-padded
          class.  The  padded  version  is  then  used  to  handle  non-linearities,  whereas the
          non-padded takes care of the rest, see demo.

   Storing datafiles
       mpi4py-fft works with regular Numpy arrays. However, since arrays in parallel  can  become
       very  large,  and  the  arrays  live  on  multiple  processors,  we  require  parallel  IO
       capabilities that goes beyond Numpys regular methods.  In the mpi4py_fft.io  module  there
       are two helper classes for dumping dataarrays to either HDF5 or NetCDF format:

          • HDF5FileNCFile

       Both  classes  have one write and one read method that stores or reads data in parallel. A
       simple example of usage is:

          from mpi4py import MPI
          import numpy as np
          from mpi4py_fft import PFFT, HDF5File, NCFile, newDistArray
          N = (128, 256, 512)
          T = PFFT(MPI.COMM_WORLD, N)
          u = newDistArray(T, forward_output=False)
          v = newDistArray(T, forward_output=False, val=2)
          u[:] = np.random.random(u.shape)
          # Store by first creating output files
          fields = {'u': [u], 'v': [v]}
          f0 = HDF5File('h5test.h5', mode='w')
          f1 = NCFile('nctest.nc', mode='w')
          f0.write(0, fields)
          f1.write(0, fields)
          v[:] = 3
          f0.write(1, fields)
          f1.write(1, fields)

       Note that we are here creating two datafiles h5test.h5 and nctest.nc, for storing in  HDF5
       or  NetCDF4  formats respectively. Normally, one would be satisfied using only one format,
       so this is only for illustration.  We  store  the  fields  u  and  v  on  three  different
       occasions, so the datafiles will contain three snapshots of each field u and v.

       Also note that an alternative and perhaps simpler approach is to just use the write method
       of each distributed array:

          u.write('h5test.h5', 'u', step=2)
          v.write('h5test.h5', 'v', step=2)
          u.write('nctest.nc', 'u', step=2)
          v.write('nctest.nc', 'v', step=2)

       The two different approaches can be used on the same output files.

       The stored dataarrays can also be retrieved later on:

          u0 = newDistArray(T, forward_output=False)
          u1 = newDistArray(T, forward_output=False)
          u0.read('h5test.h5', 'u', 0)
          u1.read('h5test.h5', 'u', 1)
          # or alternatively for netcdf
          #u0.read('nctest.nc', 'u', 0)
          #u1.read('nctest.nc', 'u', 1)

       Note that one does not have to use the same number of processors when retrieving the  data
       as when they were stored.

       It  is  also  possible  to store only parts of the, potentially large, arrays.  Any chosen
       slice may be stored, using a global view of the arrays.  It  is  possible  to  store  both
       complete fields and slices in one single call by using the following appraoch:

          f2 = HDF5File('variousfields.h5', mode='w')
          fields = {'u': [u,
                          (u, [slice(None), slice(None), 4]),
                          (u, [5, 5, slice(None)])],
                    'v': [v,
                          (v, [slice(None), 6, slice(None)])]}
          f2.write(0, fields)
          f2.write(1, fields)

       Alternatively,  one  can  use the write method of each field with the global_slice keyword
       argument:

          u.write('variousfields.h5', 'u', 2)
          u.write('variousfields.h5', 'u', 2, global_slice=[slice(None), slice(None), 4])
          u.write('variousfields.h5', 'u', 2, global_slice=[5, 5, slice(None)])
          v.write('variousfields.h5', 'v', 2)
          v.write('variousfields.h5', 'v', 2, global_slice=[slice(None), 6, slice(None)])

       In the end this will lead to an hdf5-file with groups:

          variousfields.h5/
          ├─ u/
          |  ├─ 1D/
          |  |  └─ 5_5_slice/
          |  |     ├─ 0
          |  |     ├─ 1
          |  |     └─ 3
          |  ├─ 2D/
          |  |  └─ slice_slice_4/
          |  |     ├─ 0
          |  |     ├─ 1
          |  |     └─ 2
          |  ├─ 3D/
          |  |   ├─ 0
          |  |   ├─ 1
          |  |   └─ 2
          |  └─ mesh/
          |      ├─ x0
          |      ├─ x1
          |      └─ x2
          └─ v/
             ├─ 2D/
             |  └─ slice_6_slice/
             |     ├─ 0
             |     ├─ 1
             |     └─ 2
             ├─ 3D/
             |  ├─ 0
             |  ├─ 1
             |  └─ 2
             └─ mesh/
                ├─ x0
                ├─ x1
                └─ x2

       Note that a mesh is stored along with each group of data. This mesh can be  given  in  two
       different ways when creating the datafiles:

          1. A  sequence  of  2-tuples,  where  each 2-tuple contains the (origin, length) of the
             domain along its dimension. For example, a uniform mesh  that  originates  from  the
             origin, with lengths \pi, 2\pi, 3\pi, can be given when creating the output file as:

                 f0 = HDF5File('filename.h5', domain=((0, pi), (0, 2*np.pi), (0, 3*np.pi)))

                 or, using the write method of the distributed array:

                 u.write('filename.h5', 'u', 0, domain=((0, pi), (0, 2*np.pi), (0, 3*np.pi)))

          2. A sequence of arrays giving the coordinates for each dimension. For example:

                 d = (np.arange(N[0], dtype=np.float)*1*np.pi/N[0],
                      np.arange(N[1], dtype=np.float)*2*np.pi/N[1],
                      np.arange(N[2], dtype=np.float)*2*np.pi/N[2])
                 f0 = HDF5File('filename.h5', domain=d)

       With  NetCDF4  the layout is somewhat different. For variousfields above, if we were using
       NCFile instead of HDF5File, we would get a datafile that with ncdump  -h  variousfields.nc
       would look like:

          netcdf variousfields {
          dimensions:
                  time = UNLIMITED ; // (3 currently)
                  x = 128 ;
                  y = 256 ;
                  z = 512 ;
          variables:
                  double time(time) ;
                  double x(x) ;
                  double y(y) ;
                  double z(z) ;
                  double u(time, x, y, z) ;
                  double u_slice_slice_4(time, x, y) ;
                  double u_5_5_slice(time, z) ;
                  double v(time, x, y, z) ;
                  double v_slice_6_slice(time, x, z) ;
          }

   Postprocessing
       Dataarrays  stored  to HDF5 files can be visualized using both Paraview and Visit, whereas
       NetCDF4 files can at the time of writing only be opened with Visit.

       To view the HDF5-files we first need to generate some light-weight xdmf-files that can  be
       understood  by  both  Paraview  and Visit. To generate such files, simply throw the module
       io.generate_xdmf on the HDF5-files:

          from mpi4py_fft.io import generate_xdmf
          generate_xdmf('variousfields.h5')

       This will create a number of xdmf-files, one for each group that contains 2D or 3D data:

          variousfields.xdmf
          variousfields_slice_slice_4.xdmf
          variousfields_slice_6_slice.xdmf

       These files can be opened directly in Paraview. However, note that for Visit, one  has  to
       generate the files using:

          generate_xdmf('variousfields.h5', order='visit')

       because for some reason Paraview and Visit require the mesh in the xdmf-files to be stored
       in opposite order.

   Installation
       Mpi4py-fft has a few dependencies

          • mpi4pyFFTW (serial)

          • numpycython (build dependency)

          • h5py (runtime dependency, optional)

       that are  mostly  straight-forward  to  install,  or  already  installed  in  most  Python
       environments.  The first two are usually most troublesome.  Basically, for mpi4py you need
       to have a working MPI installation, whereas FFTW is available  on  most  high  performance
       computer systems.  If you are using conda, then all you need to install a fully functional
       mpi4py-fft, with all the above dependencies, is

          conda install -c conda-forge mpifpy-fft h5py=*=mpi*

       You probably want to install into a fresh environment, though, which can be achieved with

          conda create --name mpi4py-fft -c conda-forge mpi4py-fft h5py=*=mpi*
          conda activate mpi4py-fft

       Note that this gives you mpi4py-fft with  default  settings.  This  means  that  you  will
       probably  get  the  openmpi  backend, and it is also likely that conda-forge chooses numpy
       with the mkl backend. Unfortunately, the mkl python package makes adjustments to the  FFTW
       library  and  hard to resolve bugs may arise. For this reason it is advisable to make sure
       that mkl is not installed. This can be achieved with, e.g.,

          conda create --name mpi4py-fft -c conda-forge mpi4py-fft mpich nomkl h5py=*=mpi*

       Note that the nomkl package makes sure that numpy is installed without mkl, whereas  mpich
       here chooses this backend over openmpi.

       If  you  do  not  use conda, then you need to make sure that MPI and FFTW are installed by
       some other means. You can then install any version of mpi4py-fft hosted on pypi using pip

          pip install mpi4py-fft

       whereas either one of the following will install the latest version from github

          pip install git+https://bitbucket.org/mpi4py/mpi4py-fft@master
          pip install https://bitbucket.org/mpi4py/mpi4py-fft/get/master.zip

       You can also build mpi4py-fft yourselves from the top directory, after cloning or forking

          pip install .

       or using conda-build with the recipes in folder conf/

          conda build -c conda-forge conf/
          conda create --name mpi4py-fft -c conda-forge mpi4py-fft --use-local
          conda activate mpi4py-fft

   Additional dependencies
       For storing and retrieving data you need either HDF5 or netCDF4, compiled with support for
       MPI.  HDF5  is  already  available with parallel support on conda-forge and, if it was not
       installed at the same time as mpi4py-fft, it can be installed (with the mpich backend  for
       MPI) as

          conda install -c conda-forge h5py=*=mpi_mpich_*

       A  parallel  version  of  netCDF4  cannot  be  found  on  the  conda-forge  channel, but a
       precompiled version has been made available for python 2.7, 3.6 and 3.7 on the spectralDNS
       channel, for both osx and linux

          conda install -c spectralDNS netcdf4-parallel

       Note  that  parallel  HDF5  and  NetCDF4 often are available as modules on supercomputers.
       Otherwise, see the respective packages for how to install with support for MPI.

   Test installation
       After installing (from source) it may be a good idea to run all the tests located  in  the
       tests folder. A range of tests may be run using the runtests.sh script

          conda install scipy, coverage
          cd tests/
          ./runtests.sh

       This test-suit is run automatically on every commit to github, see, e.g., .SS How to cite?

       Please cite mpi4py-fft using

          @article{jpdc_fft,
              author = {{Dalcin, Lisandro and Mortensen, Mikael and Keyes, David E}},
              year = {{2019}},
              title = {{Fast parallel multidimensional FFT using advanced MPI}},
              journal = {{Journal of Parallel and Distributed Computing}},
              doi = {10.1016/j.jpdc.2019.02.006}
          }
          @electronic{mpi4py-fft,
              author = {{Lisandro Dalcin and Mikael Mortensen}},
              title = {{mpi4py-fft}},
              url = {{https://bitbucket.org/mpi4py/mpi4py-fft}}
          }

   How to contribute?
       Mpi4py-fft  is an open source project and anyone is welcome to contribute.  An easy way to
       get started is by suggesting a new enhancement on the issue tracker. If you have  found  a
       bug,  then  either  report  this  on the issue tracker, er even better, make a fork of the
       repository, fix the bug and then create a pull request to get  the  fix  into  the  master
       branch.

   Indices and tables
       • genindex

       • modindex

       • search

AUTHOR

       Mikael Mortensen and Lisandro Dalcin

COPYRIGHT

       2020, Mikael Mortensen and Lisandro Dalcin

                                           Feb 18, 2020                             MPI4PY-FFT(1)