Storage - Duke Database Devils

Download Report

Transcript Storage - Duke Database Devils

Files and Storage: Intro
Jeff Chase
Duke University
Unix process view: data
A process has multiple channels
for data movement in and out of
the process (I/O).
I/O channels
(“file descriptors”)
stdin
stdout
tty
Process
stderr
The parent process
and parent program
set up and control
the channels for a
child (until exec).
pipe
Thread
socket
Files
Program
Files
Files
A file is a named, variable-length
sequence of data bytes that is
persistent: it exists across system
restarts, and lives until it is removed.
Unix file syscalls
fd = open(name, <options>);
write(fd, “abcdefg”, 7);
read(fd, buf, 7);
lseek(fd, offset, SEEK_SET);
close(fd);
creat(name, mode);
fd = open(name, mode, O_CREAT);
mkdir (name, mode);
rmdir (name);
unlink(name);
An offset is a byte index in a file. By default, a process reads and
writes files sequentially. Or it can seek to a particular offset.
This is called a “logical seek” because it seeks to a particular location in the file,
independent of where that data actually resides on storage (it could be anywhere).
Unix file I/O
char buf[BUFSIZE];
int fd;
Symbolic names (pathnames)
are translated through the
directory tree, starting at the
root directory (/) or process
current directory.
if ((fd = open(“../zot”, O_TRUNC | O_RDWR) == -1) {
perror(“open failed”);
File grows as process writes
exit(1);
to it  system must allocate
}
space dynamically.
while(read(0, buf, BUFSIZE)) {
if (write(fd, buf, BUFSIZE) != BUFSIZE) {
perror(“write failed”);
exit(1);
}
The file system software finds the
}
storage locations of the file’s logical
Process does not specify
blocks by indexing a per-file block map
current file offset: the
(the file’s index node or “inode”).
system remembers it.
Unix: “Everything is a file”
“Files”
regular
Afiles
Universal Set
special
Bfiles
directories
E.g., /dev/disk0s2.
The UNIX Time-Sharing System*
D. M. Ritchie and K. Thompson,1974
A symbolic name in
the file tree for a
storage volume, a
logical device.
E.g., /dev/disk0s2.
A directory/folder is
nothing more than a
file containing a list of
symbolic name
mappings (directory
entries) in some
format known to the
file system software.
Files: hierarchical name space
root directory
applications etc.
mount point
user home
directory
external media
volume or
network
storage
The file tree
A host’s file tree is the set of directories and files visible to processes
on a given host. The layout is sort of standardized, but not really.
File trees are built by grafting
subtrees from different storage
volumes or from network servers.
Each volume contains a tree of
directories and files. We can graft
it onto a directory in the file tree.
In Unix, the graft operation is
the privileged mount system call,
and each volume is a filesystem.
/
bin
ls
etc
sh
tmp
project
usr
kernel
users
packages
volume
(volume root)
mount (coveredDir, volume)
tex
emacs
coveredDir: directory pathname
volume
volume: device specifier or network volume
volume root contents become visible at pathname coveredDir
mount
point
The UNIX Time-Sharing System*
D. M. Ritchie and K. Thompson,1974
Unix file commands
• Unix has simple commands to operate on files and
directories (“file systems”: FS).
• Some just invoke one underlying syscall.
– mkdir
– rmdir
– rm (unlink)
– “ln” and “ln -s” to create names (“links”) for files
• What are the commands to create a file? Read/write a
file? Truncate a file?
Names and layers
User
view
notes in notebook file
Application
notefile: fd, byte range*
fd
bytes
block#
File System
device, block #
Disk Subsystem
surface, cylinder, sector
Add more layers as needed.
The block storage abstraction
• Read/write logical blocks of size b on a logical storage device.
• CPU (typically executing kernel code) forms buffer in memory and
issues read or write command to device queue/driver.
• Device DMAs data to/from memory buffer, then interrupts the CPU
to signal completion of each request.
• Device I/O is asynchronous: the CPU is free to do something else
while I/O in progress.
• Transfer size b may vary, but is always a multiple of some basic
block size (e.g., sector size), which is a property of the device, and
is always a power of 2.
• A logical storage device is a numbered array of these basic blocks.
• Storage blocks containing data/metadata are cached in memory
buffers while in active use: called buffer cache or block cache.
The Buffer Cache
Proc
Memory
File
cache
Ritchie and Thompson
The UNIX Time-Sharing
System, 1974
Editing Ritchie/Thompson
The system maintains a buffer cache (block cache, file
cache) to reduce the number of I/O operations.
Proc
Suppose a process makes a system call to access a
single byte of a file. UNIX determines the affected
disk block, and finds the block if it is resident in the
cache. If it is not resident, UNIX allocates a cache
buffer and reads the block into the buffer from the disk.
Then, if the op is a write, it replaces the affected byte
in the buffer. A buffer with modified data is marked
dirty: an entry is made in a list of blocks to be written.
The write call may then return. The actual write may
not be completed until a later time.
If the op is a read, it picks the requested byte out of the
buffer and returns it, leaving the block in the cache.
Memory
File
cache
Anatomy of a read
2. Enter kernel
for read syscall.
1. Compute
(user mode)
3. getBlock for maps, traverse
cached maps, getBlock for
data, and start fetch.
6. Return to
user mode.
5. Copy data from
kernel buffer to user
buffer in read.
(kernel mode)
4. sleep for I/O (stall)
CPU
seek transfer (DMA)
Disk
Time
A disk
A disk
A disk
Access time
How long to access data on disk?
– 5-15 ms on average for access to random location
– Includes seek time to move head to desired track
• Roughly linear with radial distance
– Includes rotational delay
Track
Sector
Arm
• Time for sector to rotate under head
– These times depend on drive model:
• platter width (e.g., 2.5 in vs. 3.5 in)
• rotation rate (5400 RPM vs. 15K RPM).
Cylinder
Head
• Enterprise drives use more/smaller platters spinning faster.
– These properties are mechanical and improve slowly as
technology advances over time.
Platter
Not to be tested
More than an interface — SCSI vs. ATA
D. Anderson, J. Dykes, and E. Riedel, FAST 2003
A few words about SSDs
• Technology advancing rapidly; costs dropping.
• Faster than disk, slower than DRAM.
• No seek cost. But writes require slow block erase,
and/or limited # of writes to each cell before it fails.
• How should we use them? Are they just fast/expensive
disks? Or can we use them like memory that is
persistent? Open research question.
• Trend: use them as block storage devices, and/or
combine them with HDDs to make hybrids optimized for
particular uses.
– Examples everywhere you look.
IBM Research Report
GPFS Scans 10 Billion Files in 43 Minutes
Richard F. Freitas, Joe Slember, Wayne Sawdon, Lawrence Chiu
IBM Research Division Almaden Research Center
7/22/11
The information processing…by leading business, government and scientific
organizations continues to grow at a phenomenal rate (90% CAGR).
[Compounded Annual Growth Rate]
Unfortunately, the performance of the current, commonly-used storage device -- the
disk drive -- is not keeping pace....
Recent advances in solid-state storage technology deliver significant performance
improvement and performance density improvement...
This document describes…GPFS [IBM’s parallel file system] taking 43 minutes to
process the 6.5 TBs of metadata needed for…10 Billion files. This accomplishment
combines …enhanced algorithms…with solid-state storage as the GPFS metadata
store. IBM Research once again breaks the barrier...to scale out to an unprecedented
file system size…and simplify data management tasks, such as placement, aging,
backup and replication..
HDD read bandwidth (ideal)
“spindle speed”
“Currently a high performance disk drive would have a maximum
sustained bandwidth of approximately 171 MB/s. The actual
average bandwidth would depend on the workload and the
location of data on the surface. Further, current projections do
not show much change in this over the next few years.”
IBM Research Report 2012
GPFS Scans 10 Billion Files in 43 Minutes
Enterprise disk bandwidth (2012)
2012 Seagate HDD
tomshardware.com
Max/min read bandwidth
Why does sustained
bandwidth vary by a
factor of two on the
same drive?
Areal density (storage capacity)
“The bandwidth is roughly proportional to
the linear density. So, if the growth in
linear density and track density were
equal, then one would expect the growth
rate for linear density to be the square root
of the areal density. That would make it
about 20% CAGR.”
“But, if you examine the recent history…you will see that it is
more likely to fall within the range of 10 - 15%.... Generally, the
track density has grown more quickly than the linear density.”
IBM Research Report 2011
GPFS Scans 10 Billion Files in 43 Minutes
Rotational latency
The average disk latency is ½ the rotational time of the disk drive. As you can see from its
recent history…[it] has settled down to three values 2, 3 and 4.1 milliseconds. These are ½
the inverses of 15,000, 10,000 and 7,200 revolutions per minute (RPM), respectively.
It is unlikely that there will be a disk rotational speed increase in the
near future. In fact, the 15K RPM drive and perhaps the 10K RPM
drive may disappear from the marketplace…driven by the successful
combination of SSD and slower disk drives into storage systems that
provide the same or better performance, cost and power.
Drives spin at a fixed
constant RPM. (A few
can “shift gears” to
save power, but the
gains are minimal.)
IBM Research Report 2011
GPFS Scans 10 Billion
Files in 43 Minutes
Average seek time
“The seek time is due to the
mechanical motion of the head when it
is moved from one track to another. It
is improving by about 5% CAGR. In
general, this is a mature technology
and is not likely to change dramatically
in the future. “
IBM Research Report 2011
GPFS Scans 10 Billion
Files in 43 Minutes
2012 Seagate HDD
tomshardware.com
random read access time
Disk head scheduling
FCFS: too much seeking.
What about Shortest Seek
Time First? (SSTF)
“Elevator algorithm”: sweep back
and forth, serving all requests in
one direction, then reverse.
Most of today’s drives have smart
head scheduling built in.
Memory as a cache
data
Processes access external
storage objects through file
APIs and VM abstraction. The
OS kernel manages caching of
pages/blocks in main memory.
virtual
address
spaces
data
files and
filesystems,
databases,
other
storage
objects
page/block read/write
accesses
disk and other storage
network RAM
memory
(frames)
backing storage volumes
(pages and blocks)
Memory/storage
hierarchy
Terms to know
cache index/directory
cache line/entry, associativity
cache hit/miss, hit ratio
spatial locality of reference
temporal locality of reference
eviction / replacement
write-through / writeback
dirty/clean
small
and fast
registers
(ns)
caches
L1/L2
off-core
L3
off-chip
main memory (RAM)
off-module
disk, other storage, network RAM
• In general, each layer is a cache over the layer below.
– inclusion property
• Technology trends  rapid change
• The triangle is expanding vertically  bigger gaps, more levels
big and
slow
(ms)
File Systems and Storage
Part the Second
Jeff Chase
Duke University
Storage stack
We care mostly
about this stuff.
(for now, e.g., Lab #4)
Device driver software is
a huge part of the kernel,
but we mostly ignore it.
Many storage
technologies, advancing
rapidly with time.
Databases,
Hadoop, etc.
File system API.
Generic, for use
over many kinds of
storage devices.
Standard block I/O
internal interface.
Block read/write on
numbered blocks on
each device/partition.
For kernel use only:
DMA + interrupts.
Rotational disk (HDD): cheap, mechanical, high latency.
Solid-state “disk” (SSD): low latency/power, wear issues, getting cheaper.
[Calypso]
Files as “virtual storage”
• Files have variable size.
– They grow (when a process writes more bytes past the end) and
they can shrink (e.g., see truncate syscall).
• Most files are small, but most data is in large files.
– Even though there are not so many large files, some are so large
that they hold most of the data.
– These “facts” are often true, but environments vary.
• Files can be sparse, with huge holes in the middle.
– Creat file, seek to location X, write 1 byte. How big is the file?
• Files come and go; some live long, some die young.
• How to implement diverse files on shared storage?
Using block maps
File allocation is different from heap allocation.
• Blocks allocated from a heap must be contiguous in the virtual
address space: we can’t chop them up.
• But files are accessed through e.g. read/write syscalls: the kernel
can chop them up, allocate space in pieces, and reassemble them.
• Allocate in units of fixed-size blocks, and use a block map.
• Each logical block in the object has an address (logical block
number or blockID).
• Use a block map data structure.
• Also works for other kinds of storage objects
Index map with name,
e.g., logical blockID #.
– Page tables, virtual storage volumes
– And for other kinds of maps…
– Implement in-memory cache with a hash table
Read address of the
block from map entry.
Page/block maps
Idea: use a level of indirection
through a map to assemble a
storage object from “scraps” of
storage in different locations.
The “scraps” can be fixed-size
slots: that makes allocation
easy because they are
interchangeable.
map
Example: page tables that
implement a VAS or inode
block map for a file.
http://web.mit.edu/6.033/2001/wwwdocs/ha
ndouts/naming_review.html
Representing files: inodes
• There are many many file system implementations.
• Most of them use a block map to represent each file.
• Each file is represented by a corresponding data object,
which is the root of its block map, and holds other
information about the file (the file’s “metadata”).
• In classical Unix and many other systems, this per-file
object is called an inode. (“index node”)
• The inode for a file is stored “on disk”: the OS/FS reads it
in and keeps it in memory while the file is in active use.
• When a file is modified, the OS/FS writes any changes to
its inode/maps back to the disk.
Inodes
A file’s data blocks could be “anywhere” on disk. The file’s inode maps them.
A fixed-size inode
has a fixed-size
block map.
How to represent large
files that have more
logical blocks than can
fit in the inode’s map?
attributes
block
map
inode
An inode could be “anywhere” on disk. How to find the
inode for a given file? Inodes are uniquely numbered: we
can find an inode from its number.
Once upo
n a time
/nin a l
and far
far away
,/nlived t
he wise
and sage
wizard.
data
blocks
Classical Unix inode
A classical Unix inode has a set of file attributes (below) in addition to
the root of a hierarchical block map for the file. The inode structure
size is fixed, e.g., total size is 128 bytes: 16 inodes fit in a 4KB block.
/* Metadata returned by the stat and fstat functions */
struct stat {
dev_t
st_dev;
/* device */
ino_t
st_ino;
/* inode */
mode_t
st_mode;
/* protection and file type */
nlink_t
st_nlink;
/* number of hard links */
uid_t
st_uid;
/* user ID of owner */
gid_t
st_gid;
/* group ID of owner */
dev_t
st_rdev;
/* device type (if inode device) */
off_t
st_size;
/* total size, in bytes */
unsigned long st_blksize; /* blocksize for filesystem I/O */
unsigned long st_blocks;
/* number of blocks allocated */
time_t
st_atime;
/* time of last access */
time_t
st_mtime;
/* time of last modification */
time_t
st_ctime;
/* time of last change */
};
Not to be tested
Representing Large Files
inode
Classical Unix file systems
inode == 128 bytes
inodes are packed into blocks
direct
block
map
indirect
block
Each inode has 68 bytes of attributes
and 15 block map entries that are the
root of a tree-structured block map.
Suppose block size = 8KB
12 direct block map entries: map 96KB of data.
One indirect block pointer in inode: + 16MB of data.
One double indirect pointer in inode: +2K indirects.
Maximum file size is 96KB + 16MB + (2K*16MB) + ...
The numbers on this slide are for illustration only.
double
indirect
block
indirect
blocks
Skewed tree block maps
• Inodes are the root of a tree-structured block map.
– Like multi-level hierarchical page tables, but
• These maps are skewed.
– Low branching factor at the root: just enough for small files.
– Small files are cheap: just need the inode to map it.
– Inodes for small files are small…and most files are small.
• Use indirect blocks for large files
– Requires another fetch for another level of map block
– But the shift to a high branching factor covers most large files.
• Double indirect blocks allow very large files.
• Other advantages to trees?
Post-note: what to know about maps
• What is the space overhead of the maps? Quantify.
• Understand how to lookup in a block map: logical block +
offset addressing, arithmetic to find the map entry.
• Design tradeoffs for hierarchical maps.
– Pro: less space overhead for sparse spaces.
– Con: more space overhead overall, e.g., if space is not sparse.
– Con: more complexity, multiple levels of translation.
• Skew: why better for small file files? What tradeoff?
– No need to memorize the various parameters for inode maps: concept only.
Inodes on disk
Where should inodes be stored on disk?
• They’re a good size, so we can dense-pack them into blocks. We
can find them by inode number. But where should the blocks be?
• Early Unix reserved a fixed array of inodes at the start of the disk.
– But how many inodes will we need? And don’t we want inodes
to be stored close to the file data they describe?
• Older file systems (FFS) reserve a fixed set of blocks at known
locations distributed throughout the storage volume.
• Newer file systems add a level of indirection: make a system inode
file in the volume, and store inodes in the inode file.
– That allows a variable number of inodes, and we can move them
to different locations as they’re modified.
– Originated with Berkeley’s Log Structured File System (LFS) and
NetApp’s Write Anywhere File Layout (WAFL).
Filesystem layout on disk
inode 0
bitmap file
inode 1
root directory
fixed
locations
on disk
11100010
00101101
10111101
wind: 18
0
snow: 62
0
once upo
n a time
/n in a l
10011010
00110001
00010101
allocation
bitmap file
for disk blocks
bit is set iff the
corresponding block
is in use
rain: 32
hail: 48
file
blocks
00101110
00011001
01000100
and far
far away
, lived th
inode
This is a toy example (Nachos).
A Filesystem On Disk
sector 0
sector 1
allocation
bitmap file
wind: 18
0
directory
file
11100010
00101101
10111101
snow: 62
0
once upo
n a time
/n in a l
10011010
00110001
00010101
00101110
00011001
01000100
rain: 32
hail: 48
and far
far away
, lived th
Data
A Filesystem On Disk
sector 0
sector 1
allocation
bitmap file
wind: 18
0
directory
file
11100010
00101101
10111101
snow: 62
0
once upo
n a time
/n in a l
10011010
00110001
00010101
00101110
00011001
01000100
rain: 32
hail: 48
and far
far away
, lived th
Metadata
Directories
wind: 18
0
directory
inode
snow: 62
0
rain: 32
A directory contains a set of
entries. Each directory entry is a
record mapping a symbolic name to
an inode number. The inode can
be found on disk from its number.
There can be no duplicate name
entries: the name-to-inode
mapping is a function.
hail: 48
Note: implementations
vary. Large directories
are problematic.
inode 32
Entries or free slots are typically found by a linear scan.
A creat or mkdir
operation must scan the
directory to ensure that
creates are exclusive.
Write Anywhere File Layout (WAFL)
Lab #4: DFS (“DeFiler”) buffer cache
File abstraction implemented in upper DFS layer.
All knowledge of how files are laid out on disk is at this layer.
Access underlying disk volume through buffer cache API.
Obtain buffers (dbufs), write/read to/from buffers, orchestrate I/O.
DBuffer dbuf = getBlock(blockID)
releaseBlock(dbuf)
hash
table
DBufferCache
Device I/O interface
Asynchronous I/O to/from buffers
Block read and write
Blocks numbered by blockIDs
DBuffer
read(), write()
startFetch(), startPush()
waitValid(), waitClean()
memory buffer
with header
Lab #4 DFS (“DeFiler”) interfaces
create, destroy, read, write a dfile
list dfiles
DFS
DBuffer dbuf = getBlock(blockID)
releaseBlock(dbuf)
DBufferCache
read(), write()
startFetch(), startPush()
waitValid(), waitClean()
DBuffer
ioComplete()
startRequest(dbuf, r/w)
VirtualDisk (a logical storage volume)
DBufferCache internals
HASH(blockID)
If the requested block is not resident, then getBlock
allocates a dbuf for the block and places the correct
block contents in its buffer (cache miss). If there are
no free dbufs in the cache, then we must evict some
other block from the cache and reuse its dbuf.
DBuffer dbuf = getBlock(blockID)
DBufferCache
I/O cache buffers
Each is byte[blocksize]
DBuffer
Buffer headers
DBuffer dbuf
Dbuffer (dbuf) states
DFS
A DBuffer dbuf returned by
getBlock is always associated
with exactly one block in the
disk volume. But it might or
might not be “in sync” with the
underlying disk contents.
read(…)
write(...)
startFetch(), startPush()
waitValid(), waitClean()
DBuffer
A dbuf is valid iff it has the “correct” copy of the data. A dbuf is
dirty iff it is valid and has an update (a write) that has not yet been
written to disk. A valid dbuf is clean if it is not dirty.
Your DeFiler should return only valid data to a client. That may require
you to zero the dbuf or fetch data from the disk. Your DeFiler should
ensure that all dirty data is eventually pushed to disk.
Asynchronous I/O on dbufs
Start I/O on a dbuf by posting it to a
producer/consumer queue for service by a
startFetch(),
device thread.startPush()
Client threads may wait on the dbuf
for asynchronous I/O to complete.
waitValid(), waitClean()
DBuffer
startRequest(dbuf, r/w)
Device I/O interface
Async I/O on dbufs
device threads
VirtualDisk
startFetch(), startPush()
waitValid(), waitClean()
ioComplete()
Thread upcalls dbuf
ioComplete when I/O
operation is done.
Managing files
create, destroy, read, write a dfile
list dfiles
1. Fetch blocks for data and metadata (or zero new
ones fresh) into cache buffers (dbufs).
2. Copy bytes to/from dbufs with read and write.
3. Track which data/metadata blocks are valid, and
which valid blocks are clean and which are dirty.
“inode”
for DFileID
4. Clean the dirty blocks by writing them back to the
disk with push.
DBuffer dbuf = getBlock(blockID)
releaseBlock(dbuf)
sync()
DBufferCache
DBuffer
read(), write()
startFetch(), startPush()
waitValid(), waitClean()
Filesystem layout on disk
inode 0
bitmap file
X
inode 1
root directory
X
X
11100010
00101101
10111101
Your DeFiler volume is small.
You can keep the free
block/inode maps in memory.
You don’t need metadata
structures on disk for that. But
you have to scan the disk to
rebuild the in-memory
structures on initialization.
DeFiler must be
able to find all
valid inodes on
disk.
X
0
once upo
n a time
/n in a l
rain: 32
hail: 48
file
blocks
and far
far away
, lived th
inode
DeFiler has no directories.
You just need to keep track
of which DFileIDs are
currently valid, and return a
list.
Disk layout: the easy way
DeFiler must be able
to find all valid
inodes on disk.
Given a list of valid inodes,
you can determine which
inodes and blocks are free and
which are in use.
once upo
n a time
/n in a l
file
blocks
and far
far away
, lived th
inode
Unix file naming: hard links
directory A
A Unix file may have multiple names.
Each directory entry naming the file is
called a hard link.
Each inode contains a reference count
showing how many hard links name it.
directory B
0
rain: 32
wind: 18
0
hail: 48
sleet: 48
inode link
count = 2
inode 48
link system call
link (existing name, new name)
create a new name for an existing file
increment inode link count
unlink system call (“remove”)
unlink(name)
destroy directory entry
decrement inode link count
if count == 0 and file is not in active use
free blocks (recursively) and on-disk inode
Illustrates: garbage collection by reference counting.
Unix file naming: soft links
A symbolic or “soft” link is a file whose contents is the pathname of
another file. They are useful to customize the name tree, and also can
be confusing and error-prone.
symlink system call
symlink (existing name, new name)
allocate a new file (inode) with type symlink
directory A
directory B
initialize file contents with existing name
0
wind: 18
create directory entry for new file with new name
0
rain: 32
sleet: 67
hail: 48
inode link
count = 1
../A/hail/0
inode 48
See command “ln –s”.
The target of the soft link
may be removed at any
time, leaving a dangling
reference.
inode 67
How should the kernel
handle recursive soft links?
Unix file naming: links
usr
ln -s /usr/Marty/bar bar
Lynn
creat foo
unlink foo
foo
Marty
creat bar
ln
/usr/Lynn/foo
bar
unlink
bar
bar
Concepts
• Reference counting and reclamation
• Redirection/indirection
• Dangling reference
• Binding time (create time vs. resolve time)
• Referential integrity
Post-note: symbolic name maps
• Hierarchy for symbolic names (directory hierarchy):
– Multiple naming contexts, possibly under control of different
owners. E.g., each directory is a separate naming context.
– Avoids naming conflicts when people reuse the same names.
– Pathname lookup by descent through the hierarchy from some
starting point, e.g., root (/) or current directory.
– Build the name space by subtree grafting: mounts.
– Accommodates different directory implementations per-subtree.
• E.g., modern Unix mixes FS implementations through Virtual File
System (VFS) layer.
– Scales to very large name spaces.
• Note: Domain Name Service (DNS) is the same!
– www.cs.duke.edu “==“ /edu/duke/cs/www
Virtual memory
Memory
0:
1:
Page Table
Virtual
Addresses
0:
1:
Physical
Addresses
CPU
P-1:
N-1:
Disk
VMs (or segments) are storage objects described
by maps. A page table is just a block map of one
or more VM segments in memory. The hardware
hides the indirection from the threads that are
executing within that VM.
CMU 15-213
Cartoon view of a page table
Each process/VAS has
its own page table.
Virtual addresses are
translated relative to
the current page table.
process page table (map)
PFN 0
PFN 1
PFN i
In this example, each
VPN j maps to PFN j, but
in practice any physical
frame may be used for
any virtual page.
PFN i
+
offset
VPN #i
offset
user virtual address
physical memory
page frames
Virtual page: a logical block in a segment.
VPN: Virtual Page Number (a logical block number).
Page frame: a physical block in machine memory.
PFN: Page Frame Number (a block pointer).
PTE: Page Table Entry (an entry in the block map).
The maps are themselves
stored in memory; a
protected CPU register
holds a pointer to the
current map.
Example: Windows/IA32
•
Two-level block map (page table) structure reduces the space
overhead for block maps in sparse virtual address spaces.
–
•
Many process address spaces are small: e.g., a page or two of text, a
page or two of stack, a page or two of heap.
Windows provides a simple example of a hierarchical page table:
–
Each address space has a page directory (“PDIR”)
–
The PDIR is one page: 4K bytes, 1024 4-byte entries (PTEs)
–
Each PDIR entry points to a map page, which MS calls a “page table”
–
Each map page (“page table”) is one page with 1024 PTEs
–
Each PTE maps one 4K virtual page of the address space
–
Therefore each map page (page table) maps 4MB of VM: 1024*4K
–
Therefore one PDIR maps a 4GB address space, max 4MB of tables
–
Load PDIR base address into a register to activate the VAS
Page table structure for a process on
Windows on IA 32 architecture
Step 2. Index page table with PT2
Step 1. Index PDIR with PT1
virtual address
32 bits
Two-level page table
32-bit virtual address
Two 10-bit page table index fields (PT1, PT2)
(10 bits represents index values 0-1023)
[from Tanenbaum]
Virtual Address Translation
12
Example: typical 32-bit
architecture with 4KB pages.
0
Virtual address translation maps a
virtual page number (VPN) to a
physical page frame number
(PFN): the rest is easy.
VPN
offset
address
translation
Deliver exception to
OS if translation is not
valid and accessible in
requested mode.
physical address
{
PFN
+
offset
More pictures
• We did not discuss these last three pictures to help
understand name mapping structures.
• COW: one advantage of page/block maps is that it
becomes easy to clone (logical copy) a block space.
– Copy a storage object P to make a new object C. P could be a
file, segment, volume, or virtual address space (for fork!).
– Copy the map P: make a new map C referencing the same
blocks. The map copy is cheap: no need to copy the data itself.
– Since a clone is a copy, any changes (writes) to P after the
clone should not affect C, and vice versa.
– Use a lazy copy or copy-on-write (COW). Intercept writes
(how?) and copy the affected block before executing the write.
Copy on write
Physical
memory
Parent
memory
Child
memory
What happens if parent writes to a page?
Landon Cox
Copy on write
Physical
memory
Parent
memory
Child
memory
Have to create a copy of pre-write page for
the child.
Landon Cox
Virtual Addressing: Under the Hood
MMU
start
here
probe
page table
load
TLB
yes
miss
probe
TLB
load
TLB
access
physical
memory
access
valid?
hit
zero-fill
no
raise
exception
OS
no (first reference)
fetch
from disk yes
page on
disk?
allocate
frame
legal
reference
page
fault?
signal
process
illegal
reference
How to monitor page reference events/frequency along the fast path?
Replacement policy: file systems
• File systems often use a variant of LRU.
– A file system sees every block access (through syscall API), so it
can do full LRU: move block to tail of LRU list on each access.
• Sequential files have a cache wiping effect with LRU.
– Most file systems detect sequential access and prefer eviction of
blocks from the same file, e.g., using MRU.
– That also prevents any one file/object from consuming more than
its “fair share” of the cache.
VM systems
• VM memory management is similar to file systems.
– Page caching in physical memory frames
– Unified with file block caching in most systems
– Virtual address space is a collection of regions/segments, which
may be considered as “objects” similar to files.
• Only it’s different.
– Mapped by page tables
– VM system software does not see most references, since they
are accelerated by Memory Management Unit hardware.
– Requires a sampling approximation for page replacement.
– All data goes away on a system failure: no write atomicity.
VM page replacement
• Try to guess the working set of pages in active use for each VAS.
• To determine if a page is being used, arrange for MMU to notify OS
on next use.
– E.g., reference bit, or disable read access to trigger a fault.
• Sample pages systematically to approximate LRU: e.g., CLOCK
algorithm, or FIFO-with-Second-Chance (FIFO-2C)
Why “logical” devices/volumes?
The block storage abstraction is an abstraction! We can
implement block storage in a wide variety of ways.
• Partition a block space on some physical device into multiple
smaller logical devices (logical volumes).
• Concatenate devices to form a larger logical volume.
• Add software and indirection (block maps) to map a space of
logical blocks to a dynamic mix of underlying devices and/or servers.
• Servers and/or devices can implement block storage service over a
network: network disk, network storage, …
– Storage Area Network (SAN) or iSCSI (Internet SCSI).
– Network-Attached Storage (NAS) generally refers to a network
file system abstraction, built above block storage.
• Add another level of indirection! Storage virtualization.
NAS, SAN, and all that
[rtcmagazine.com]
File Systems and Storage
Day Three
Performance and Reliability
Jeff Chase
Duke University
Storage software stack
System call interface for files
Related VM operations (mmap)
Generic FS layer (e.g., VFS)
+“file system drivers” for specific FS
getBlock(blockID)
releaseBlock(…)
startFetch(), startPush()
waitValid(), waitClean()
I/O buffering / page cache layer
ioComplete()
startRequest(…)
VirtualDisk layer
Storage system performance
• How to get good storage performance?
– Build better disks: new technology, SSD hybrids.
– Gang disks together into arrays (RAID logical devices).
– Smart disk head scheduling (when there is a pool of pending
requests to choose from).
– Smarter caching: better victim selection policies
– Asynchronous I/O: prefetching, read ahead, “write behind”
– Location, location, location: smart block placement
• It’s a big part of the technology of storage systems.
Memory as a cache
data
Processes access external
storage objects through file
APIs and VM abstraction. The
OS kernel manages caching of
pages/blocks in main memory.
virtual
address
spaces
data
files and
filesystems,
databases,
other
storage
objects
page/block read/write
accesses
disk and other storage
network RAM
memory
(frames)
backing storage volumes
(pages and blocks)
Performance of logical storage
• Let us always remember that a logical storage volume can be
implemented in all kinds of wild ways: storage virtualization.
• Even “simple” devices have complex mapping/translation internally.
– E.g., Flash Translation Layer to spread write load across SSD device.
– E.g., disk electronics automatically hide bad blocks on the platter.
• So: it is hard to generalize about performance behavior.
– “All generalizations are false.”
• How can we build higher-level storage abstractions (like file
systems or databases) above block storage?
• How can they use the device(s) efficiently? How much do they
need to know about storage performance properties?
– E.g., “seeks waste time”
Block placement and layout
• One key assumption: “seeks waste time”.
– Blocks whose addresses (logical block numbers) are close
together are cheaper to access together.
– “Sequentialize!”
• Location, location, location:
– Place data on disk carefully to keep related items close together
(smart block allocation).
– Use larger b (larger blocks, clustering, extents, etc.)
– Smaller s (placement / ordering, sequential access, logging, etc.)
Effective bandwidth
Seeks are overhead: “wasted effort”. It is a cost s that the device
imposes to get to the data. It is not actually transferring data.
This graph is obvious. It applies to so many things in computer
systems and in life.
b/(sB+b)
1
Effective bandwidth is
efficiency or goodput
What percentage of the
time is the busy resource
(the disk head) doing useful
work, i.e., transferring data?
100%
Spindle bandwidth B
b/B
Transfer size b
s (seek)
Effective bandwidth
Effective bandwidth or bandwidth utilization is the share or
percentage of potential bandwidth that is actually delivered. E.g., what
percentage of time is the disk actually transferring data, vs. seeking etc.?
Define
b Block size
B Raw disk bandwidth (“spindle speed”)
s Average access (seek+rotation) delay per block I/O
Then
Transfer time per block = b/B
I/O completion time per block = s + (b/B)
Delivered bandwidth for I/O request stream = bytes/time = b/(s + (b/B))
Bandwidth wasted per I/O: sB
So
Effective bandwidth: bandwidth utilization/efficiency (%): b/(sB + b)
[bytes transferred over the “byte time slots” consumed for the transfer]
Effective bandwidth by access time
b/(sB+b)
1
100%
Spindle bandwidth B (90 MB/s
b=256K
b=64K
b=4K
access time s
5ms
Bigger is better. Other things being equal, effective bandwidth is higher
when access costs can be amortized over larger transfers. High access
cost is the reason we use tapes primarily for backup! As B grows and s is
unchanged, disks are looking more and more like tapes! (Jim Gray)
Prefetching for high read throughput
• Read-ahead (prefetching)
– Fetch blocks into the cache in expectation that they will be used.
– Requires prediction. Common for sequential access.
1. Detect access pattern.
2. Start prefetching
Reduce I/O stalls
Sequential read-ahead
• Prediction is easy for sequential access. “Most files
are read and written sequentially.”
• Read-ahead also helps reduce seeks by reading
larger chunks if data is laid out sequentially on disk.
App requests block n
App requests block n+1
n+1
n
n+2
System prefetches block n+2
System prefetches block n+3
Building better file systems
• The 1990s was a period of experimentation with new strategies for
high-performance file system design.
• The new file systems generally used the FFS mechanisms and
data structures, but changed the policies for block allocation.
– Block allocation policy: where to place new data (or modified old data)
on the storage volume? Which block number to choose?
– “File system design is 99% block allocation.” - Larry McVoy
– Example: Group large-file data into big contiguous chunks called
clusters or extents that can be read or written as a unit (larger b).
[McVoy91] and [Smith/Seltzer96]
– Example: Write modified data and metadata wherever convenient to
minimize seeking: e.g., “log-structured” file systems (LFS)
[Rosenblum91] and NetApp’s WAFL [Hitz95]. Note: requires a level of
indirection so the FS can write each version of an inode to a different
location on the disk. (See WAFL’s inode file.)
Fast File System (FFS)
• Fast File System (FFS) [McKusick81] is the historical, canonical
Unix file system that actually works.
• In the old days (1970s-1980s), file systems delivered only 10% of
the available disk bandwidth, even on the old disks.
• FFS extended the classic 1970s Unix file system design with a new
focus on performance in the Berkeley Unix release (BSD: 1982).
– Multiple block sizes: use small blocks called frags in small files,
to reduce internal fragmentation.
– Smart block allocation that pays attention to disk locality by
placing data/metadata in zones called cylinder groups.
• FFS was still lousy, but it laid the groundwork for development of
high-performance file systems over the next 20 years.
FFS block allocation policy
• FFS partitions space on a disk into logical regions as a zone of
locality. When it allocates a block, it chooses the region carefully.
– A cylinder group is a region of contiguously numbered disk blocks that
are believed to “probably” reside on a group of adjacent tracks on the
disk. The idea is that seeks within a CG are “short”.
– Every block on the disk resides in exactly one CG. Blocks in the same
CG are “close together”. Blocks in different CGs are “far apart”.
• Policy: Place “related” data in the same CG whenever possible.
• Policy: Smear large files across CGs, so they don’t fill up a CG.
• Policy: Reserve space for inodes in each CG, so inodes can be
close to directory entries that reference them.
• Policy: Place maps (inodes or indirect blocks) in the
same CG as the data blocks they reference.
• You can see the impact of these policies in the plots!
A quick peek at sequential write
in BSD Unix “FFS” (Circa 2001)
note sequential
block allocation
physical
disk
sector
write
write stall
read
sync command
(typed to shell)
pushes indirect
blocks to disk
sync
read next
block of
free space
bitmap (??)
time in milliseconds
Sequential writes: a closer look
physical
disk
sector
16 MB in one second
(one indirect block worth)
140 ms
delay for
cylinder seek
etc. (???)
time in milliseconds
Push indirect blocks synchronously.
longer delay
for head movement
to push indirect
blocks
write
write stall
Small-File Create Storm
50 MB
inodes and
file contents
(localized allocation)
physical
disk
sector
note synchronous
writes for some
metadata
sync
sync
delayed-write
metadata
sync
write
write stall
time in milliseconds
When to write? When to sync?
Some metadata blocks
are written
synchronously: the
system waits for the disk
writes to complete before
continuing.
Sync is a system call: it
pushes all dirty data out
of the cache, and waits
for the writes to complete.
Why?
Some metadata blocks are written delayed
(writeback): they sit dirty in the cache and then are
pushed out to disk at a later convenient time (e.g.,
when the cache has “too much” dirty data).
Writes in the FS buffer cache
• Delayed writes
– Partial-block writes may be left dirty in the cache. The “push” to
disk is deferred in expectation of later writes to the same block.
• Write-behind
– Dirty blocks file are pushed to disk asynchronously; the write
syscall may return before the disk write is complete.
– May lose data! Be sure you know the failure semantics of the
file systems you use in your life. A classic UNIX file system may
discard any subset of recent writes on failure.
• Fsync syscall pushes dirty blocks and waits for them.
– Fsync is for use by applications that really want to know their
data is “safe”. Good file systems implicitly fsync-on-close.
Metadata updates and recovery
• Metadata updates may incur extra seek overhead.
– E.g., extending a file requires writes to the inode, direct
and/or indirect blocks, cylinder group bit maps and
summaries, and the file block itself.
• Metadata items are often updated repeatedly, so
delayed writes help.
• But delayed writes incur a higher risk of file system
corruption in a crash.
– Suppose the metadata structure is in an inconsistent state
after a crash, and can’t be repaired? Then what?
– If you lose your metadata, you are dead in the water.
Safety of metadata
• How to protect integrity of the metadata structures?
– Metadata is a complex linked data structure, e.g., a tree.
– Must be “well-formed” after a crash/restart, even if writes are lost.
– …or, must be possible to restore metadata to a consistent state
with a scrub (file system check or “fsck”) on restart after a crash.
once upo
n a time
/n in a l
wind: 18
dir inode
and far
far away
, lived th
file blocks
0
snow: 62
file inode
0
rain: 32
hail: 48
dir entries
Atomic updates: the recovery problem
The safe metadata update problem in file systems is a simplified form of
the atomic update and recovery problem for databases.
•
We want to make a group of related updates to a complex linked data structure,
e.g., to create a new file. The updates could be all over the disk.
•
But we could crash at any time, e.g., in the middle of the group of updates.
•
We need some way to do atomic commit: either all of the updates in each
group complete, or none of them do. And we want it to be fast.
•
The concern is similar to concurrency control: we don’t want software to “see”
an inconsistent state that violates structural invariants.
once upo
n a time
/n in a l
wind: 18
0
dir inode
and far
far away
, lived th
file blocks
snow: 62
file inode
0
rain: 32
hail: 48
dir entries
Failures, Commits, Atomicity
• Hard state is state (data) that a service needs to
function correctly. Soft state is non-critical.
• What guarantees does the system offer about the hard
state if the system fails?
– Durability
• Did my writes commit, i.e., are they saved? Can I get it back?
– Failure atomicity
• Can a set of writes “partly commit”?
– Recoverability and Corruption
• Can we recover a previous state?
• Is the state consistent?
• Is the metadata well-formed?
Disk write behavior (cartoon version)
• Disk may reorder pending writes.
– Limited ordering support (“do X before Y”).
– Can enforce ordering by writing X synchronously: wait for write
of X to complete before issuing Y.
• Writes at sector grain are atomic (512 bytes?).
• Writes of larger blocks may fail “in the middle”.
• Disk may itself have a writeback cache.
– Even “old” writes may be lost.
– (The cache can be disabled.)
Atomic commit: techniques
We consider three three techniques for atomic commit
and recovery, in the context of file systems.
• Option 1: careful write ordering, with scrub on recovery
• Option 2: logging/journaling (also used in databases)
• Option 3: shadowing (e.g., WAFL)
Option 1
Metadata write ordering
A common approach to safety of file system metadata:
• Order essential metadata writes carefully.
– Various techniques to enforce a partial order of writes to the disk,
i.e., ensure that write A completes before write B begins.
• Maintain invariants! E.g., avoid dangling references.
– Never recycle a structure (block or inode) before zeroing all
pointers to it (truncate, unlink, rmdir).
– Never point to a new structure before it has been initialized.
E.g., sync inode on creat before filling directory entry, and sync a
new block before writing the block map.
• Traverse the metadata tree to rebuild derived structures.
– Post-crash scrub can rebuild/repair other structures e.g., free
block bitmaps, free inode bitmaps.
Option 2
Logging/Journaling
• Logging is widely used for database systems, and for
metadata writes in “journaling” file systems.
• Key idea: record updates in a sequential log file as they
are made.
– Log records are written to the log synchronously and
sequentially: no seeks, and preserves temporal ordering.
– Each log record write is atomic: each log record is trailed by a
marker (e.g., checksum) that says “this log record is complete”.
• Commit each group g of related writes atomically by
writing a single commit record to the log: “commit g”.
– To recover: scan the log in order, reapply (“replay”) committed
updates and/or cancel or roll back updates from any group g that
did not commit before the crash.
Recoverable Data with a Log
volatile memory (e.g., buffer cache)
Your program (or file system
or database software)
executes transactions that
read or change the data
structures in memory.
Your data structures
in memory
Log transaction
events as they occur
Push a checkpoint or
snapshot to disk
periodically
snapshot
log
After a failure, replay the log into the last snapshot to recover the data structure.
Transactions
Database systems and other systems use a programming
construct called atomic transactions to represent a group
of related reads/writes, often on different data items.
BEGIN T1
read X
read Y
…
write X
END
Transactions commit
atomically in a serial order.
BEGIN T2
read X
write Y
…
write X
END
Transactions: logging
Commit
WriteN
Write1
Begin transaction
Append info about modifications to a log
Append “commit” to log to end x-action
Write new data to normal database
Single-sector write commits x-action (3)
Begin
1.
2.
3.
4.

…
Transaction Complete
Invariant: append new data to log before applying to DB
Called “write-ahead logging”
Transactions: logging
Commit
WriteN
…
Write1
Begin transaction
Append info about modifications to a log
Append “commit” to log to end x-action
Write new data to normal database
Single-sector write commits x-action (3)
Begin
1.
2.
3.
4.

What if we crash here (between 3,4)?
On reboot, reapply committed updates in log order.
Transactions: logging
WriteN
…
Write1
Begin transaction
Append info about modifications to a log
Append “commit” to log to end x-action
Write new data to normal database
Single-sector write commits x-action (3)
Begin
1.
2.
3.
4.

What if we crash here?
On reboot, discard uncommitted updates.
Anatomy of a Log
• A log is a sequence of records
(entries) on recoverable storage.
• Each entry is associated with
some transaction T.
• Create log entries for T as T
executes, to record progress of T.
• Atomic operations:
– Append/write entry to log
– Truncate older entries up to time t
(old)
Log
Sequence
Number
(LSN)
...
LSN 11
XID 18
LSN 12
XID 18
Transaction
ID (XID)
LSN 13
XID 19
commit
record
LSN 14
XID 18
commit
– Read/scan entries
• Log writes are atomic and durable,
and complete detectably in order.
(new)
Using a Log
• Log entries for T record the writes
by T (or operations in T).
– Redo logging
• To recover, read the checkpoint
and replay committed log entries.
(old)
Log
Sequence
Number
(LSN)
– Skip the records of uncommitted Ts
• No T can be allowed to affect the
checkpoint until T commits.
– Technique: write-ahead logging
LSN 11
XID 18
LSN 12
XID 18
Transaction
ID (XID)
LSN 13
XID 19
– “Redo” by reissuing writes or
reinvoking the methods.
– Redo in order (old to new)
...
commit
record
LSN 14
XID 18
commit
(new)
Managing a Log
• On checkpoint, truncate the log
– No longer need the entries to recover
• Checkpoint how often? Tradeoff:
– Checkpoints are expensive, BUT
– Long logs take up space
(old)
Log
Sequence
Number
(LSN)
– Is it safe to redo/replay records whose
effect is already in the checkpoint?
– Checkpoint “between” transactions,
so checkpoint is a consistent state.
– Lots of approaches
LSN 11
XID 18
LSN 12
XID 18
Transaction
ID (XID)
LSN 13
XID 19
– Long logs increase recovery time
• Checkpoint+truncate is “atomic”
...
commit
record
LSN 14
XID 18
commit
(new)
File Systems and Storage
Day Four
Filers and Service Performance
Jeff Chase
Duke University
Atomic updates: the recovery problem
The safe metadata update problem in file systems is a simplified form of
the atomic update and recovery problem for databases.
•
We want to make a group of related updates to a complex linked data structure,
e.g., to create a new file. The updates could be all over the disk.
•
But we could crash at any time, e.g., in the middle of the group of updates.
•
We need some way to do atomic commit: either all of the updates in each
group complete, or none of them do. And we want it to be fast.
•
The concern is similar to concurrency control: we don’t want software to “see”
an inconsistent state that violates structural invariants.
once upo
n a time
/n in a l
wind: 18
0
dir inode
and far
far away
, lived th
file blocks
snow: 62
file inode
0
rain: 32
hail: 48
dir entries
Atomic commit: techniques
We consider three three techniques for atomic commit
and recovery, in the context of file systems.
• Option 1: careful write ordering, with scrub on recovery
• Option 2: logging/journaling (also used in databases)
• Option 3: shadowing (e.g., WAFL)
Shadowing
1. starting point
2. write new blocks to disk
modify purple/grey blocks prepare new block map
Shadowing is a basic technique for atomic
commit and recovery. It is used in WAFL.
3. overwrite block map
(atomic commit)
and free old blocks
(optional)
Just to spell it out: if the system crashes before step 3, then the update fails, but the
previous version is still intact. To abort the failed update we just need to free any
blocks written in step 2. Step 3 completes the update: it replaces the old map with the
new. Because it is a single disk write, the system cannot fail “in the middle”: it either
completes or it does not: it is atomic. Once it is complete, the new data is safe.
On-disk metadata structures
Write Anywhere File Layout (WAFL)
WAFL and Writes
• Any modified data/metadata can go anywhere on the disk.
– The WAFL metadata structure assures this: every piece of metadata
is linked in a tree rooted in the root pointer.
• An arbitrary stream of updates can be installed atomically.
– Retain the old copy: “no overwrite”
– Switch to new copy with a single write to the root (shadowing).
• WAFL’s design naturally maintains multiple point-in-time
consistent snapshots of each file volume.
– Old copy lives on as a point-in-time snapshot.
WAFL Snapshots
The snapshot mechanism is used for user-accessible
snapshots and for transient consistency points.
WAFL’s on-disk structure (high level)
Root Inode
Metadata
File Data Blocks
NetApp Confidential - Limited Use
118
Another Look
WAFL and Performance
• Write the new updated copies wherever and whenever it is
“convenient” (fast).
• NetApp filers are designed to be “scalable”: add modules to make
the system more powerful.
•
E.g., add more disks (RAID).
“Filers”
• Network-attached (IP)
• RAID appliance
• Multiple protocols
– iSCSI, NFS, CIFS
• Admin interfaces
• Flexible configuration
• Lots of virtualization: dynamic volumes
• Volume cloning, mirroring, snapshots, etc.
• NetApp technology leader since 1994 (WAFL)
Network File System (NFS)
Remote Procedure Call (RPC)
External Data Representation (XDR)
[ucla.edu]
Remote Procedure Call (RPC)
• NFS was an early popular application of RPC.
– “RPC is a canonical structuring paradigm for client/server
request/response services.”
– Used in .NET, Android, RMI, distributed component frameworks
client
server
Auto-generate this code
from API spec (IDL).
“glue”
sockets
Humans focus on
getting this code right.
sockets
This code is “canned”,
independent of the
specific application.
WAFL and the disk system
• WAFL generates a continuous stream of large-chunk
contiguous writes to the disk system.
– WAFL does not overwrite the old copy of a modified structure: it
can write a new copy anywhere. So it gathers them together.
• Large writes minimize seek overhead and deliver the full
bandwidth of the disk.
• WAFL gets excellent performance by/when using many
disks in tandem (“RAID”)…
• …and writing the chunks in interleaved fashion across
the disks (“striping”).
• Old copies of the data and metadata survive on the disk
and are accessible through point-in-time “snapshots”.
Incremental Scalability
• Scalability: what does it really mean?
How do we measure or
validate claims of scalability?
not
scalable
scalable
cost
marginal
cost of capacity
No hockey
sticks!
capacity
Scaling and bottlenecks
Scale up by adding capacity incrementally?
• “Just add bricks/blades/units/elements/cores”...but that presumes
we can parallelize the workload.
• Vertically: identify functional stages, and execute different stages
on different units (or “tiers”).
• Horizontally: spread requests/work across multiple units.
– Or partition the data and spread the chunks across the elements, e.g.,
for parallel storage or parallel computing.
• Load must be evenly distributed, or else some element or stage
saturates first (bottleneck).
A bottleneck limits throughput
and/or may increase response
time for some class of requests.
Work
Benchmarks and performance
• Benchmarks enable standardized comparison under
controlled conditions.
– Compare “apples to apples”, avoid misrepresentation e.g., by
vendor selected benchmarks (“atrocities in the marketplace” –
Jeff Mogul).
• They embody some specific set of workload
assumptions.
• Subject a system to a selected workload, and measure
its performance.
• Server/service metrics:
– Throughput: request/sec at peak (saturation)
– Response time: t(response) – t(request)
http://www.spec.org/sfs2008/
Ideal throughput: cartoon version
throughput == arrival rate
The server is not saturated: it
completes requests at the rate
requests are submitted.
throughput == peak rate
The server is saturated. It can’t
go any faster, no matter how
many requests are submitted.
Ideal throughput
Response
rate
(throughput)
i.e., request
completion
rate
saturation
peak rate
Request arrival rate (offered load)
This graph shows
throughput (e.g., of a
server) as a function
of offered load. It is
idealized: your
mileage may vary.
Throughput: reality
Thrashing, also called congestion collapse
Real servers/devices often have some pathological behaviors at
saturation. E.g., they abort requests after investing work in them
(thrashing), which wastes work, reducing throughput.
Response
rate
(throughput)
i.e., request
completion
rate
delivered
throughput
(“goodput”)
saturation
peak rate
Request arrival rate (offered load)
Illustration only
Saturation behavior is
highly sensitive to
implementation
choices and quality.
Improving throughput
1. Make the service center faster. (“scale up”)
–
Upgrade the hardware, spend more $$$
2. Reduce the work required per request (D).
–
More/smarter caching, code path optimizations, use smarter
disk layout.
3. Add service centers, expand capacity. (“scale out”)
–
RAIDs, blades, clusters, elastic provisioning
–
N centers improves throughput by a factor of N: iff we can
partition the workload evenly across the centers!
–
Note: the math is different for multiple service centers, and
there are various ways to distribute work among them, but we
can “squint” and model a balanced aggregate roughly as a
single service center: the cartoon graphs still work.
This graph shows how certain design alternatives under study impact a
server’s throughput. The alternatives reduce per-request work
(overhead) and/or improve load balancing. (This is a graph from a random
research paper: the design alternatives themselves are not important to us.)
saturation
Measured
throughput
(“goodput”)
Higher
numbers
are better.
Note how throughput
degrades in overload
on this system.
Offered load (request/sec)
RAID 0
Striping
•
•
•
•
•
•
Sequential throughput?
Random throughput?
Random latency?
Read vs. write?
MTTF/MTBF?
Cost per GB?
Fujitsu
RAID 1
Mirroring
•
•
•
•
•
•
Sequential throughput?
Random throughput?
Random latency?
Read vs. write?
MTTF/MTBF?
Cost per GB?
Fujitsu
Building a better
disk: RAID 5
• Redundant Array of
Independent Disks
• Striping for high throughput
for pipelined reads.
• Data redundancy: parity
• Enables recovery from one
disk failure
• RAID5 distributes parity:
no“hot spot” for random
writes
• Market standard
Fujitsu
The remaining slides were not covered in class.
They deal with the response time curve and why
always “bends up”. They are not to be tested.
Scaling and response time
In the real world we don’t want to saturate our systems.
We want systems to be responsive, and saturated systems aren’t responsive.
Instead, characterize max
request rate λmax this way:
1. Define a response time
objective: maximum
acceptable response time
(Rmax): a simple form of
Service Level Objective
(SLO).
2. Increase λ until system
response time surpasses
Rmax : that is λmax.
λmax
Rmax
λ
[graphic from IBM.com]
Queuing Theory for Busy People
wait here in queue
offered load
request stream @
arrival rate λ
Process for mean
service demand D
“M/M/1” Service Center
• Big Assumptions (at least for this summary)
– Single service center (e.g., one core)
– Queue is First-Come-First-Served (FIFO, FCFS).
– Independent request arrivals at mean rate λ (poisson arrivals).
– Requests have independent service demands at the center.
– i.e., arrival interval (1/λ) and service demand (D) are
exponentially distributed (noted as “M”) around some mean.
– These assumptions are rarely true for real systems, but they give
a good “back of napkin” understanding of queue behavior.
Response time (R)
R == D
The server is idle. The
response time of a request is
just the time to service the
request (do requested work).
R = D + queuing delay
As the server approaches
saturation, the queue of waiting
request grows without bound.
(We will see why in a moment.)
saturation (U = 1)
Rmax
Average
response
time R
U
R
saturation
D
λmax
Request arrival rate (offered load)
Illustration only
Saturation behavior is
highly sensitive to
implementation
choices and quality.
The same picture, only different
Queuing delay is
proportional to:
Response time,
determined by:
rho is “load factor”
λ/λmax
= r/rmax
= utilization
“stretch factor”
R/D
(normalized
response time)
(Max request load)
“saturation” λmax
Offered load
(requests/sec)
also called lambda (λ)
Little’s Law
• For an unsaturated queue in steady state, mean
response time R and mean queue length N are
governed by:
– Little’s Law: N = λR
Why?
• Suppose a task T is in the system for R time units.
• During that time:
– λR new tasks arrive (on average)
– N tasks depart (all the tasks ahead of T, on average).
• But in steady state, the flow in balances flow out.
– Note: this means that throughput X = λ in steady state.
Inverse Idle Time “Law”
Service center saturates as 1/ λ
approaches D: small increases in λ
cause large increases in the expected
response time R.
R
U
1(100%)
Little’s Law gives response time R = D/(1 - U).
Intuitively, each task T’s response time R = D + DN.
Substituting λR for N: R = D + D λR
Substituting U for λD: R = D + UR
R - UR = D  R(1 - U) = D  R = D/(1 - U)
Why Little’s Law is important
1. Intuitive understanding of FCFS queue behavior.
Compute response time from demand parameters (λ, D).
Compute N: how much storage is needed for the queue.
2. Notion of a saturated service center.
Response times rise rapidly with load and are unbounded.
At 50% utilization, a 10% increase in load increases R by 10%.
At 90% utilization, a 10% increase in load increases R by 10x.
3. Basis for predicting performance of queuing networks.
Cheap and easy “back of napkin” estimates of system
performance based on observed behavior and proposed
changes, e.g., capacity planning, “what if” questions.
Guides intuition even in scenarios where the assumptions of the
theory are not met.
Utilization: cartoon version
U = XD
X = throughput
D = service demand, i.e., how
much time/work to complete
each request.
1 == 100%
Utilization
(also called
load factor)
U = 1 = 100%
The server is saturated. It has
no spare capacity. It is busy all
the time.
saturated
saturation
peak rate
Request arrival rate (offered load)
This graph shows
utilization (e.g., of a
server) as a function
of offered load. It is
idealized: each
request works for D
time units on a single
service center (e.g., a
single CPU core).
Utilization
• What is the probability that the center is busy?
– Answer: some number between 0 and 1.
• What percentage of the time is the center busy?
– Answer: some number between 0 and 100
• These are interchangeable: called utilization U
• The probability that the service center is idle is 1-U.
The Utilization Law
• If the center is not saturated then:
– U = λD = (arrivals/T) * service demand
• Reminder: that’s a rough average estimate for a mix
of independent request arrivals with average service
demand D.
• If you actually measure utilization at the device, it
may vary from this estimate.
– But not by much.