1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
ΓΓΡΑΦ ΊΣ
πδεκδθωθά
ηααέ ηαμ
χ δεΪ η
ηΪμ
ΨΆΞΙΜΟ
Πκζδ δεά Απκλλά κυ
Φυ δεά ωηα δ έωθ:
Θ ωλέα ωθ χκλ υθ εαδ β
φτ β βμ πλαΰηα δεσ β Ϊμ
ηαμ
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ · 31
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
ε ηίλέκυ 2017
1ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
SHARE
Labels
Astronomy
Big Bang
cosmology
dimensions
nature
Physics
space
Γα
α,
α
α
α α
φυ
α απ
υ
α
π
α
α π
α
υ
α απ
α π α α
π
α
α π α απ
αυ
υ
π
π
α
α
υ
.
α
φυ
.
string theory
universe
α
α υπ α
α απ υ υ
α
ππ
υ
α.
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
βή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
πα
απ
υ
υ
π υ α
α
α
π
αυ
α. Κ
α
α
,
α απ
π
π
α
πα
α
α
υ π
, υ α
φυ
φα
υ
π
α α
αα α α
α,
π
π υ
υ α
α
π υ
απ υ
α υ π
υ
α α, α π α
α
α
υ
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
α
υ
φ
φυ
α
α. Ό α
π
π υ π
π
.
α ,π π
α α α
α
α
υ
φ
υ
α απ
υ
α
υ
. Αυ
απ υ π
φ
α
α απ
α απ
υ π
α α α
ππ
υ π
υπ
α α
απ υ υ
π
υ
απ
α υ
α
α
π
α α
ππ
φυ
π
α
π
π α α
φυ
υ
α
φα
υ π
α
α
α
.
φ
π
υ
.
γή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
α
α
α
α
υ
α
α
α
π
π α α
φ
α
φ
α
α
α
α.
υ,
α
π
απ
,
α
α
π
,
π
.
α
α α α
α
π
φα
α
π
α φα α
α
α α
α π υ
α
α π
αυ
αα α
υ
υ
α απ αυ
υ . αυ
α α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
υ
α
υ .
α απ α π
α
α. υ
α
α
α
.
α π
φυ
π
α
α
α
α αα φ
π υ
α
π α α
α α α
α
φα
α
π
υ π υ α
υ α π υα
π
υ
α
υ
π
π υ
π
α
α
υ π
α π
απ υ
α
π
α
υ
π
α
α
απ υα
φυ
υ
υ
π
α α
α.
υ
α α
π
, π
υ
α
α
υ
α,
α
υ .
4ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
Ό α αφ
απ
απ α π
α
Garry McKinnon π α α π
α
,
π
α α α
α
π
υ ,
απ
α.
α
φα α,
π α
,
α απ
α
απ αυ
α
π υ
α α
φ
α α
υ
α
α
π
α
π υ
π
α
πα
α .Κ
π α α
αυ
π
α
α
α α
π
α α
Gary McKinnon
α α
α
π
α
φ
α
α
π
α. Α
φ
α
α,
α
α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
υ
π α α
υ αφ
α
α π α α
Α α
α
α.
π α α
φυ
π
Leumeria,
π
π
π
π
α
υ
φυ
αυ
υ φυ
υπ υ
5ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
α . Απ α α υ φ
π υ
π υ
π
,
α
φ υ
απ υ
α
α π υ απ
υπ
α
),
α,
, υ
,α
α αυ
α
α
α
π
π
α α,
'αυ
π α α
π
α α
α
π υ
α
α
π υ
α απ
π
φ
π
α α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
α ,α
π α α
α
υπ α
π α α
υ
α
α
,
π
υ α
α
φ
φυ
φυ
α,
υ π
π α α
π
φ
υ
φυ
α
αα
φυ
υ
α
Γ
π ,(
υ
υ
υ
α
φ α.
α.
υ
α
α
α φ
α
.
π
π
α
π α
π υ
,
π
υ
π
αυ
υ
α
υ , φυ
αφυ
Φα
π
α α α α
π
φυ
α
π
υ
α α
α
π
α.
υ
αυ
α
.Α
υ
6ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
α α
α
φ
. π
α
α
α απ
φυ
π
π
α ,α
π α α
π
α
π α
α
α
α α
π
φ
α
π
υ
αυ
α
.Α
φυ
π
α
α ,α
π α α
α
α.
α
α α
υ
υ
φυ
υ
υ
α,
α
α α α
'αυ
α
φ
απ
φυ
α
.Α
.
π
π
π α α
υ
π α α
α,
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
υ
α α
α α
Θεωρία γρα
α
α
υπ α
αυ
α
π α α
α α
π
π
φ
απ
π
π α α
α απ
α
υ
α.
υ
'αυ
α α α
α
φ
α,
π α α
α
α
.
ώ
ιή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
Απ
π
υ
π
υ,
π
α
Η
α
απ υπ π
ππ
,
α
υ
π α α
α α α
α α
π
πα
αυ
φα
π υ υ
π α α
α απ
.
αφ
α
π
υ
φ
φ
α,
,α
α
α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
π υ
,
α
.
α
α
φυ
α
.
π
υ
α
,
π
α
υ π
α π α α
υ
α
α α. π
α απ
α
Αυ
φ
φυ
α α
α.
α
πα α
α
α α
απ αυ
απ
α.Η υ
απ
υ
αυ
α απ
α π
α
υ .
υ α
α απ
α π υ
α
απ
α
π α α
φυ
π
α π α α
υ
α.
απ π
α
κή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
π
υ
α
Αυ
π υ
υ
υ
α (α
π α
α)
π υ
υ
π α
α π
φ
υ
π
,
α
α;
υ
υ
π
α
φυ
α ,α
α
υ
,
, α α
α
α
υ
αφ
α
α
π . Απ
υ
α
α
π
αφ
α π υ πα α
π α υ
α
α
α
α
φα
π
αυ
Αυ
υ
φ
α
α
π α
φ
α
α.
π α α
απ
φυ
α α
υ
α υπ
π
α π α α
πα
α,
α π
φ
υ
α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
α
αυ
υ
.
α
υπ
α
πα
α
α, π
α π
α
.
α, α π α
9ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
υ
α
φ
υ
απ υ
φ
υ
π
.Η
α
α
,
α
φ
υ
π υ
α
α
α.
α
π
π υ π
.
απ
,
α υ
α
;
α
π
υ α
α α
α
υ
α
α ,π
α
υ .
α απα
α
π α
απ αυ
π υ
π
υ
υ
α
Η
α.
φ
π α α
,α
αα
α
α
απ υπ
απ
α
α
απ
αφ
α
α
απ
α υπ
π
α π
.Η
α π
π
υ
α
α
υ π
α
α
α
υ
υ
α
α α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
π
α υ
α
υ
α
φυ
α
υ ,
α
1ίή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
π α α
α, π π
α
α
υ
π
αυ
απ
α
α
α
Α
υ
υ
π
α
α
α π
α
υ
α
α
απ
α.
υ α
α α
α.
φυ
α α υπ
π
α α α
α
, α
π
α α
π α α
υ
.
α α α
α
υ
π υπ
.
φα
α α, α υ α
υ
π υ α
αφ
φυ
π υ
α
υ
;
α
.
υ
π
α
.
υ
πα
υ
υ α
α . Κα π
π α α
πα α
υπ
π υ
αυ
π
α
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
α α
υ
αυ
α
α
υ
α
α
π α
α
11ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
υ
α . Αυ
π
π
α
α α α
α
α
π υ
α.
The video below shows that scientists are still trying to
determine weather or not it would be physically possible
to go through a wormhole. At the same time, I believe our
black budget world has already developed the technology
to do so. Like I mentioned earlier, this is why I believe
there is a group on the planet that operates at a
completely different level of understanding about science
and technology, while the mainstream world is left to
ponder what is already known.
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
1βή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
Michio Kaku Explains String Theory
Sources:
http://www.damtp.cam.ac.uk/research/gr/public/qg_ss
.html
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
1γή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
http://plus.maths.org/content/string-theory-newtoneinstein-and-beyond
http://www.dummies.com/how-to/content/the-basicelements-of-string-theory.html
ΣΙΚΈΣ : ΑΣΡΟΝΟΜ Α , BIG BANG , ΚΟΜΟΛΟΓ Α ,
ΣΗ Φ Η , ΦΤΙΚ , Χ ΡΟ , Θ ΩΡ Α ΣΩΝ ΧΟΡ
ΙΑΣΆ Ι ,
Ν , ΣΟ ΜΠΑΝ
Μ ΡΊ ΙΟ
χσζδα
Enter your comment...
ΠλΫπ δ θα δαίΪ
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
14ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
έθαδ αζζκ απσμ πκυ
πλκ παγ έ θα
πδεκδθωθά δ η β Γβ;
Πβΰά ωθ ζ υ αέωθ
«αζζκ απυθ» βηΪ ωθ
Ϋχ δ απκηαελτθ δ κυμ
α λκθσηκυμ
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ ·
27
ε ηίλέκυ 2017
Μ ΡΊ ΙΟ
Ι
ΗΜΟΊ ΤΗ ΧΟΛΊΟΤ
ΆΣ Π ΡΙΌΣ Ρ
Οδ πδ άηκθ μ
απκεαζτπ κυθ πλυ μ
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
15ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
παλα βλβ δεΫμ
απκ έι δμ σ δ κ τηπαθ
ηαμ ά αθ εΪπκ Ϋθα
κζσΰλαηηα
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ ·
26
ε ηίλέκυ 2017
Μ ΡΊ ΙΟ
Ι
1 ΧΌΛΙΟ
ΆΣ Π ΡΙΌΣ Ρ
Σκ ζδεσ σλδκ αχτ β αμ
εαδ β πέ λα β κυ
χλσθκυ
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ · 04 Ικυζέκυ 2017
Μ ΡΊ ΙΟ
Ι
ΗΜΟΊ ΤΗ ΧΟΛΊΟΤ
ΆΣ Π ΡΙΌΣ Ρ
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
16ή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
Οδ φυ δεκέ Ϋζθκυθ
ωηα έ δα φω σμ κ
παλ ζγσθ,
απκ δεθτκθ αμ σ δ κ
αιέ δ κ χλσθκ έθαδ
φδε σ!
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ ·
30
ε ηίλέκυ 2017
Μ ΡΊ ΙΟ
Ι
ΗΜΟΊ ΤΗ ΧΟΛΊΟΤ
ΆΣ Π ΡΙΌΣ Ρ
Ο πδ άηκθαμ Brian Cox
ζΫ δ σ δ ΰθωλέα δ ΰδα έ θ
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
1ιή1κ
1ή1ήβί1κ
Φυ δεά ωηα δ έωθμ Θ ωλέα ωθ χολ υθ εαδ η φτ η ημ πλαΰηα δεσ η άμ ηαμ
ίλάεαη αζζκ απκτμ
Μ Πλαΰηα δεσ β α πΫλα απσ βθ τζβ ·
28
ε ηίλέκυ 2017
Μ ΡΊ ΙΟ
Ι
3 ΧΌΛΙ
ΆΣ Π ΡΙΌΣ Ρ
Φοί λσ SPACE
Powered by Blogger
Θέματα από enot-poloskun
httpμήήwww.realitybeyondmatter.comήβί1ιή1βήparticle-physics-string-theory-and.html
1κή1κ
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
λγλο
α α
α πυ
π
υ
, Μα γαέομ . Broome, Casey R. Myers, Andrew G. White & Timothy C. Ralph
Martin Ringbauer
Nature Communications 5 , αλδγησμ αλξ έου: 4145
Λάοβ: 07 Νο ηβλέου 2013
πο
(2014)
Δβηο δ υηΫθο online: 19 Ιουθέου 2014
doi : 10.1038 / ncomms5145
Κα βΪ
ε σ: 16 Μα ου 2014
βθ παλαποηπά
Θ πλβ δεά φυ δεά
Αφηρη
η
υ
α π
φυ
,
αυ
α
- απα α
α π
α. Η
αυ
α α α
α
α
υ π
φ
υ αυ
α απ υπ
π
υ, α
π υ
αφ
υ α,
α
υ
α α
α
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
α α
α
α π
π
π
α
α
α
υ
υ
α
α α
α
υ
α
α π
ξσζδα
,
α α
α υ α πα
α
α
.
π
π
απ
α
α
α .Ε
qubit π υ α
α
α α
α
α
α απ υ
π υ
α
α α
ο νέο χέ ιο. φά
υ
, π
,
πα
π
α υ
Βλέπ
α
, αφ
φ
α
υΑ
.
υ
α π
υ
α πα
α π υ
α
υ
timelike. Αυ
πα α
φα
α πα
πα α
π
π
α
απ
α φ
π
α α
π
απ
.
α
α
α απ
υ
α,
π υ
αυ
1ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
ππ
απ
α
π
α
α
α
α
π
α
α
α
απ
φ
,
υ
αα
.
Ει αγωγ
Μαπ υ
υ
α π υ
α
π
π α π
υ
π
π υ π
φ υ
φα
α υ π
π
α
π
α
α
α
α
α
π
π
π
α
απ
8
α
πα
α πα
α
πα α α
α
Deutsch CTCs
υ
π
,
υ
,
α
α
υπ
υ π
7
πα
αφ
1991
α
α
α
CTC.
υ CTCs
υπ
α α, π
α,
α
π
α
α
υ
υ
π
α φ
α
. Επ
,
πα α
πα
απ
π
α π
α π υ
7 , 15
π
υ
π
φ
πα
α
α
α
πα α α
υ
α
απ υ
α π
α
α
υ
9
α α
απα
α
α
π υ
υ
10 , 11
α
α
α α
NP-π
π
αφ α,
α, π
π
α
υ Heisenberg 12 .
α
π
π
α
,
α α
α
π υ
π
α α
φα α.
α απ υ
α
απ
.
υ
α
υ
π υ
α-
α π υ
π
υ α
υΑ
α.
π
Η
α α
α
π
υ π
π
υ Hawking 5 ,
π
,
- υπ
α α
αφ
α
α
α
1,2,3
υπ
υ
υ
υ παππ
α
υ Deutsch 6
α
(CTCs) - α
υ Novikov 4
υ
π υ α α απ αυ
α
π
πα
α
α
α
α π
- π
αυ
α
υ φυ
α
α α
α
α
φ
α
π
α
α CTC
υ φα
α
α
α π
α
,
υ
α υ π
13 , 14
.
αυ
,
α υ π
.
α α
α
π υ πα
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
π
υ π
π
φ
π
α
π
υ π
φ
α
α
2ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
υ
.Α
απ
16 , 17 , 18
α
20 , 21 , 22 , 23
π
π
φα
υ
α
π
π
α α
α
α α
π
α
. Αυ
αφ
α α
π
π
φ
υ
α
α φα
υ
α α
υ
α
φυ
υ
α
φ
αυ - υ π
,
Απο ε
ο ο
α
α
α
α
υ
α
φ
α
α
π
π
α
α
α
| ψ >, α
α
qubit CTC-
υ
Deutsch
υ6.
απ υπ
Bacon 8 α
υ
α
υ π
π
U
υ
π
α
απ
φ
υ υ
,π
απ
α
π
π α απ
αφ
απ
υ υ
π υ
.
υ Brun et al. 9
υπ
υα
υ π
π
απ
υπ
,
α
α
απ
υα .7
α
α
π υπ
α π
α
,
α
α α
υ
π
π
α
α
α
α α
19 ,
φ
αυ α
α
α α
α
υ
π
υ υα
υ α
π
α α
α
α
απ
Deutsch CTC.
υ
π υ α,
α
απ CTC π
υ
π
υ α
traversing qubit
φ
α
α
αυ
α α
φ
α α
υ
α
α
υ .
.
α α
ο Deutsch
While there has been some recent success on alternative models of CTCs based
on post-selection23,24,25, we focus on the most prominent model for describing
quantum mechanics in the presence of CTCs, introduced by Deutsch6. Here a
quantum state |ψLj interacts unitarily with an older version of itself (Fig. 1). With
the inclusion of an additional swap gate, this can equivalently be treated as a
two-qubit system, where a chronology-respecting qubit interacts with a qubit
ρCTC trapped in a CTC. The quantum state of ρCTC in this picture is determined by
Deutsch’s consistency relation:
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
3ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Figure 1: Model of a quantum state |ψ
〉 interacting with an older version of
itself.
This situation can equivalently be interpreted as a chronology-respecting qubit interacting with a qubit
trapped in a CTC. The CTC in general consists of a causal worldline with its past and future ends
connected via a wormhole (indicated by black triangles).
where U′ is the unitary U followed by a SWAP gate (Fig. 1). This condition ensures
physical consistency—in the sense that the quantum state may not change inside
the wormhole—and gives rise to the nonlinear evolution of the quantum state |
ψLj. The state after this evolution is consequently given by ρOUT=Tr2[U′(|ψLjLJψ|
⊗ρCTC)U′†]. The illustration in Fig. 1 further shows that the requirement of
physical consistency forces ρCTC to adapt to any changes in the surroundings,
such as a different interaction unitary U or input state |ψLj. While equation (1) is
formulated in terms of a pure input state |ψLj, it can be directly generalized to
mixed inputs7.
Simulating CTCs
Our experimental simulation of a qubit in the (pure) state |ψLj traversing a CTC
relies on the circuit diagram shown in Fig. 2a). A combination of single-qubit
unitary gates before and after a controlled-Z gate allows for the implementation
of a large set of controlled-unitary gates U. Using polarization-encoded single
photons, arbitrary single-qubit unitaries can be realized using a combination of
quarter-wave (QWP) and half-wave plates (HWP); additional SWAP gates before or
after U are implemented as a physical mode swap. The controlled-Z gate is based
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
4ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
on non-classical (Hong-Ou-Mandel) interference of two single photons at a
single partially polarizing beam splitter (PPBS) that has different transmittivities
V=1/3
for vertical (V) and
H=1
for horizontal (H) polarization26—a more detailed
description of the implementation of the gate can be found in ref. 27.
Conditioned on post-selection, it induces a π phase shift when the two
interfering single-photon modes are vertically polarized, such that |VVLj
→−|VVLj with respect to all other input states.
Figure 2: Experimental details.
(a) The circuit diagram for a general unitary interaction U between the state |ψLj and the CTC system. (b)
The specific choice of unitary in the demonstration of the (i) nonlinear evolution and (ii) perfect
discrimination of non-orthogonal states. (c) Experimental setup for case (ii). Two single photons,
generated via spontaneous parametric down-conversion in a nonlinear β-barium-borate crystal, are
coupled into two optical fibres (FC) and injected into the optical circuit. Arbitrary polarization states are
prepared using a Glan–Taylor polarizer (POL), a quarter-wave (QWP) and a half-wave plate (HWP). Nonclassical interference occurs at the central partially polarizing beam splitter (PPBS) with reflectivities
H=0 and V=2/3. Two avalanche photo-diodes (APDs) detect the single photons at the outputs. The
states |ΨLj are chosen in the x–z plane of the Bloch sphere, parameterized by ϕ, and CUxz is the
corresponding controlled unitary, characterized by the angle xz. The SWAP gate was realized via
relabelling of the input modes.
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
ηή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
One of the key features of a CTC is the inherently nonlinear evolution that a
quantum state |ψLj undergoes when traversing it. This is a result of Deutsch’s
consistency relation, which makes ρCTC dependent on the input state |ψLj. In
order to simulate this nonlinear behavior using linear quantum mechanics we
make use of the effective nonlinearity obtained from feeding extra information
into the system. In our case we use the classical information about the
preparation of the state |ψLj and the unitary U to prepare the CTC qubit in the
appropriate state ρCTC as required by the consistency relation equation (1). After
the evolution we perform full quantum state tomography on the CTC qubit in
order to verify that the consistency relation is satisfied.
Nonlinear evolution
As a first experiment we investigate the nonlinearity by considering a Deutsch
CTC with a CNOT interaction followed by a SWAP gate as illustrated in Fig. 2b(i).
This circuit is well known for the specific form of nonlinear evolution:
which has been shown to have important implications for complexity theory,
allowing for the solution of NP-complete problems with polynomial resources8.
According to Deutsch’s consistency relation (equation (1)) the state of the CTC
qubit for this interaction is given by
We investigate the nonlinear behaviour experimentally for 14 different quantum
states
, with
and a variety of
phases φ∈{0, 2π}, where the locally available information ϕ and φ is used to
prepare ρCTC. In standard (linear) quantum mechanics no unitary evolution can
introduce additional distinguishability between quantum states. To illustrate the
nonlinearity in the system we thus employ two different distinguishability
measures: the trace distance
, where
, and a single
projective measurement with outcomes ‘+’ and ‘−’:
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
θή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
While the metric is a commonly used distance measure, it does not have an
operational interpretation and requires full quantum state tomography in order
to be calculated experimentally. The measure in contrast is easily understood as
the probability of obtaining different outcomes in minimum-error discrimination
of the two states using a single projective measurement on each system. The
operational interpretation and significance of is discussed in more detail in the
Supplementary Note 1. Both and are calculated between the state |ψLj and the
fixed reference state |HLj after being evolved through the circuit shown in Fig.
2b(i). The results are plotted and compared to standard quantum mechanics in
Fig. 3. If the state |ψLj is not known then, based only on the knowledge of the
reference state |HLj and the evolution in equation (2) it is natural and optimal to
use the measure with a σz-measurement.
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
ιή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Figure 3: Nonlinear evolution in a Deutsch CTC with SWAP.CNOT interaction.
Both the trace distance and the σz-based distinguishability measure (equal to within experimental error
in this case) of the evolved states ρOUT a er the interaction with the CTC are shown as yellow diamonds.
The blue circles (red squares) represent the measure () between the input states |ψLj and |HLj in the
case of standard quantum mechanics. Note that due to the phase independence of the evolution in
equation (2), states that only di er by a phase collapse to a single data point. Crucially, the metric does
not capture the e ect of the nonlinearity, while does, indicated by the red-shaded region. Error bars
obtained from a Monte Carlo routine simulating the Poissonian counting statistics are too small to be
visible on the scale of this plot. Inset: The dashed black lines with decreasing thickness represent
theoretical expectations for and from 2, 3, 4 and 5 iterations of the circuit.
We observe enhanced distinguishability for all states with an initial trace
distance to |HLj smaller than
(that is,
), as clearly demonstrated by
the σz-based measure, see Fig. 3. Note, however, that this advantage over
standard quantum mechanics is not captured by the metric (ρ1, ρ2) unless the
nonlinearity is amplified by iterating the circuit on the respective output at least
three times, see the inset of Fig. 3. This shows that the nonlinearity is not
directly related to the distance between two quantum states. By testing states
with various polar angles for each azimuthal angle on the Bloch sphere, we
confirm that any phase information is erased during the evolution and that the
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
κή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
evolved state ρOUT is indeed independent of φ, up to experimental error. We
further confirm, with an average quantum state fidelity of F=0.998(2) between
the input and output states of ρCTC in equation (3), that the consistency relation
(1) is satisfied for all tested scenarios.
Non-orthogonal state discrimination
While it is the crucial feature, nonlinear state evolution is not unique to the
SWAP.CNOT
interaction, but rather a central property of all non-trivial CTC
interactions. Similar circuits have been found to allow for perfect
distinguishability of non-orthogonal quantum states9, leading to discomforting
possibilities such as breaking of quantum cryptography9, perfect cloning of
quantum states10,11 and violation of Heisenberg’s uncertainty principle12. In
particular it has been shown that a set
of N distinct quantum states in a
space of dimension N can be perfectly distinguished using an N-dimensional CTC
system. The algorithm proposed by Brun et al.9 relies on an initial SWAP operation
between the input and the CTC system, followed by a series of
controlled
unitary operations, transforming the input states to an orthogonal set, which can
then be distinguished.
In our simulation of this effect we consider the qubit case N=2, which
consequently would require two controlled unitary operations between the input
state and the CTC system. We note, however, that without loss of generality the
set of states to be discriminated can be rotated to the x–z plane of the Bloch
sphere, such that |ψ0Lj=|HLj and
for some angle ϕ.
In this case, the first controlled unitary is the identity operation , while the
second performs a controlled rotation of |ψ1Lj to |VLj as illustrated in Fig. 4a). In
detail, the gate CUxz applies a π rotation to the target qubit conditional on the
state of the control qubit, about an axis in the x–z plane defined by the angle
For the optimal choice
xz.
the gate rotates the state |ψ1Lj to |VLj,
orthogonal to |ψ0Lj, enabling perfect distinguishability by means of a projective
σz measurement (see Fig. 4a).
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
λή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Figure 4: Bloch-sphere evolution of states traversing a CTC.
In the case of (a) local state preparation, the state |ψ0Lj=|HLj (blue) is una ected by CUxz, while |Ψ1Lj
(green) undergoes a π rotation about the axis defined by xz. The axis is chosen as
such that
, which can then be perfectly distinguished from |ψ0Lj. (b) For non-local
preparation of the initial states and the same choice of xz the controlled unitary maps both initial states
to the maximally mixed state
. The probability of distinguishing the two states
is therefore 1/2—as good as randomly guessing.
In practice the gate CUxz is decomposed into a controlled-Z gate between
appropriate single-qubit rotations, defining the axis
xz.
The latter are realized by
half-wave plates before and after the PPBS, set to an angle of
xz/4
with respect
to their optic axis (see Fig. 2c):
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1ίή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Note that relation (1) requires that ρCTC=|HLjLJH|, whenever the input state is
|HLj, independent of the gate CUxz. Crucially, this consistency relation ensures
that any physical CTC system adapts to changes in ϕ and
xz,
parameterizing the
input state and gate, respectively. In our simulation these two parameters are
used to prepare the corresponding state ρCTC, as shown in Fig. 2c.
In a valid experimental simulation the input and output states ρCTC have to
match, that is, ρCTC has to satisfy relation (1). This has been verified for all
following experiments with an average quantum state fidelity of =0.996(7).
In the experiment, we prepared near-pure quantum states directly on single
photons using a Glan–Taylor polarizer followed by a combination of a HWP and a
QWP. We simulated CTC-aided perfect discrimination of non-orthogonal states
for 32 distinct quantum states |ψ1Lj with ϕ∈[0, 2π). For each state we
implemented CUxz with the optimal choice of
. Furthermore, we
tested the ability of this system to distinguish the set {|ψ0Lj, |ψ1Lj} given nonoptimal combinations of ϕ and
xz.
For this we either chose ϕ=3π/2 and varied
the gate over the full range of
, or chose CUxz as a controlled
Hadamard (optimal for ϕ=3π/2) and varied the state |ψ1Lj over the full range of
ϕ∈[0, 2π). The output state is characterized by quantum state tomography, which
provides sufficient data to obtain for arbitrary measurement directions as well as
for the calculation of the trace distance.
Figure 5a illustrates the observed distinguishability for the above experiments
and compares it to the expectation from standard quantum mechanics. In the
latter case the measure is maximized by choosing the optimal projective
measurement, based on the available information about the states |ψ0Lj and |ψ1Lj.
Crucially, the optimized L is directly related to the trace distance as
and therefore captures the same qualitative picture, without the requirement for
full quantum state tomography. In the CTC case a σz-measurement is chosen,
which is optimal when
based on the knowledge of
. Otherwise further optimization is possible
xz
(see Supplementary Note 2 and Supplementary
Fig. 1 for more details). Furthermore, we note that the above scenario can also be
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
11ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
interpreted in a state-identification rather than state-discrimination picture,
which is discussed in more detail in Supplementary Note 3 and Supplementary
Fig. 2.
Figure 5: Experimental results.
Probability of state discrimination for (a) locally prepared and (b) non-locally prepared states |ψ0Lj
=|HLj and
as measured by . The surface represents the
theoretically predicted probability depending on the state and gate parameters ϕ and xz, respectively.
Solid, red (open, blue) data points indicate better (worse) performance than standard quantum
mechanics. (c) Cross-sectional views of the combined plots a and b reveal the rich structure in the
dependencies on the initial parameters for (top) a fixed state (ϕ=3π/2) and (bottom) a fixed gate
( xz=π/4). Here red squares (yellow diamonds) correspond to the CTC case with local (non-local)
preparation and blue circles represent standard quantum mechanics. Error bars obtained from a Monte
Carlo routine simulating the Poissonian counting statistics are too small to be visible on the scale of this
plot.
Local versus non-local state preparation
Owing to the inherent nonlinearity in our simulated system, care must be taken
when describing mixed input states ρin. In particular a distinction between
proper and improper mixtures can arise, which is unobservable in standard
(linear) quantum mechanics28. This ambiguity is resolved in ref. 7 by requiring the
consistency condition to act shot-by-shot—that is, independently in every run of
the experiment—on the reduced density operator of the input state. For proper
mixtures this means that ρin is always taken as a pure state, albeit a different one
shot-by-shot. For improper mixtures in contrast, ρin will always be mixed. A
similar, but much more subtle and fascinating feature that has received less
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
12ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
attention in the literature so far occurs with respect to preparation of pure
states29. While in standard quantum mechanics a pure state prepared directly
(locally) on a single qubit is equivalent to one that has been prepared non-locally
through space-like separated post-selection of an entangled resource state, this
is not true under the influence of a CTC. The origin of this effect is not the
nonlinear evolution, but rather the local absence of classical information about
the post-selection outcome. The role of locally available classical information in
entanglement-based preparation schemes is a matter of current debate and still
to be clarified.
A possible resource state for alternatively preparing |ψ0Lj and |ψ1Lj could be of
the form
, where projection of the first qubit onto
the state |0Lj and |1Lj leaves the second qubit in the state |ψ0Lj and |ψ1Lj,
respectively. From the point of view of ρCTC, however, there exists no information
about the outcome of this projective measurement. Hence it ‘sees’ and adapts to
the mixed state
. The state of the CTC qubit
is therefore different for local and non-local preparation. If this was not the case,
it would enable superluminal signalling, which is in conflict with relativity29.
Figure 4b) illustrates the evolution induced by CUxz, when the input states |HLj
and |ψ1Lj are prepared using an entangled resource | Lj, rather than directly.
The results of the previously discussed distinguishability experiments for this
case are shown in Fig. 5b). In Fig. 5c) they are compared to the case of local
preparation and to standard quantum mechanics for a fixed input state and a
fixed gate, respectively. Again, consistency of our simulation is ensured by a
quantum state fidelity of =0.9996(3) between the input and output states of ρCTC.
In our simulation we find that the CTC system can indeed achieve perfect
distinguishability of the (directly prepared) states |ψ0Lj and |ψ1Lj even for
arbitrarily close states if the appropriate gate is implemented (see Fig. 5a).
Furthermore, we show that the advantage over standard quantum mechanics
persists for a wide range of non-optimal gate–state combinations, outside of
which, however, the CTC system performs worse (blue points). Notably, we find
that for non-locally prepared input states CTC-assisted state discrimination
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
13ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
never performs better than random guessing—a probability of
(as shown in
Fig. 5b). The predictions for standard quantum mechanics, in contrast are
independent of the way the states |ψ0Lj and |ψ1Lj are prepared.
Decoherence
We further investigated the effect of two important decoherence mechanisms on
the simulated CTC system (shown in Fig. 2a). The first is a single-qubit
depolarizing channel acting on the input state |ψLj, which can be modelled as
where (σx, σy, σz) are the three Pauli matrices and p∈[0, 1] quantifies the amount
of decoherence.
The second form of decoherence concerns the controlled unitary CUxz and is
described as
where ε∈[0, 1] is the probability of the gate to fail, describing the amount of
decoherence that is present. For ε=0 the gate acts as an ideal controlled rotation
CUxz, while it performs the identity operation for ε=1.
We tested the robustness of the state-discrimination circuit in Fig. 2b(ii) against
both forms of decoherence. For this test we chose CUxz as a controlled
Hadamard (that is,
xz=π/4)
and the initial states |ψ0Lj=|HLj and
(that is, ϕ=3π/2). Fig. 6 shows the distinguishability of the
evolved states as a function of both decoherence mechanisms over the whole
range of parameters p∈[0, 1] and ε∈[0, 1]. Note that the decoherence channel in
equation (7) does not have an analogue in the standard quantum mechanics case
(that is, without a CTC); hence only the channel in equation (6) is considered for
comparison. It is further naturally assumed that the experimenter has no
knowledge of the specific details of the decoherence and therefore implements
the optimal measurements for the decoherence-free case. The physical validity
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
14ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
of the simulation is ensured by the consistency of ρCTC across the boundary of
the wormhole with an average fidelity of =0.997(4).
Figure 6: State discrimination as a function of gate and qubit decoherence
for locally prepared states.
Here ε quantifies the decoherence of the unitary interaction CUxz (with xz=π/4), which has no analogue
in the standard quantum mechanics case, and p the single-qubit depolarization of the input qubits |HLj
and |ψ1Lj (with ϕ=3π/2). The system demonstrates robustness against both forms of decoherence and
the CTC advantage persists up to
and
, respectively. The semi-transparent blue
surface represents the optimum in standard quantum mechanics. Error bars obtained from a Monte
Carlo routine simulating the Poissonian counting statistics are too small to be visible on the scale of this
plot.
It is worth noting that the interpretation of decoherence effects in the circuit in
Fig. 2a) is very different from the linear scenario without a CTC. In the case of
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1ηή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
single-qubit depolarization the initially pure input state becomes mixed. In
contrast to the linear case now an important distinction has to be made with
respect to the origin of the decoherence. If it results from an interaction with the
environment, which is the case considered here, then ρCTC ‘sees’ an improper
mixture and adjusts to the mixed density matrix of the input state. If, however,
the origin of the mixture is classical fluctuations in the preparation apparatus,
then shot-by-shot pure states enter the circuit and the consistency relation
holds accordingly shot-by-shot, resulting in a proper mixture at the output. This
suggests that in the presence of a CTC it would be possible to identify the origin
of the decoherence in an experimental setup.
Furthermore, careful analysis of the decoherence of the unitary gate CUxz reveals
parallels to the effects seen in non-local state preparation. The decoherence is
assumed to arise from non-local coupling to the environment. Again, due to a
lack of classical knowledge of the outcome of an eventual measurement of the
environment, ρCTC ‘sees’ the mixed process in equation (7) in every run of the
experiment. In the case of full decoherence the distinguishability is reduced to
0.5 as in standard quantum mechanics. The differences between local and nonlocal decoherence in their interpretation and effect is one of the key insights
from our simulation.
Discussion
Quantum simulation is a versatile and powerful tool for investigating quantum
systems that are hard or even impossible to access in practice20. Although no
CTCs have been discovered to date, quantum simulation nonetheless enables us
to study their unique properties and behaviour. Here we simulated the
immediate adaption of ρCTC to changes in the CTC’s environment and in
particular the effect of different forms of decoherence. We also show that the
nonlinearity inherent in the system is in fact not uniform (as shown in Fig. 3),
suggesting that nonlinear effects only become apparent in certain scenarios and
for a specific set of measurements.
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1θή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Moreover, we find intriguing differences with respect to nominally equivalent
ways of pure state preparation. Although acknowledged in ref. 29, this feature
has not been further investigated in the present literature. Importantly, this
effect arises due to consistency with relativity, in contrast to the similar effect
for mixed quantum states discussed earlier, which is a direct result of the
nonlinearity of the system7.
Our study of the Deutsch model provides insights into the role of causal
structures and nonlinearities in quantum mechanics, which is essential for an
eventual reconciliation with general relativity.
Additional information
How to cite this article: Ringbauer, M. et al. Experimental simulation of closed
timelike curves. Nat. Commun. 5:4145 doi: 10.1038/ncomms5145 (2014).
References
1.
Morris, M. S., Thorne, K. S. & Yurtsever, U. Wormholes, time machines, and
the weak energy condition. Phys. Rev. Lett. 61, 1446–1449 (1988).
2.
Morris, M. S. & Thorne, K. S. Wormholes in spacetime and their use for
interstellar travel: a tool for teaching general relativity. Am. J. Phys. 56, 395–
412 (1988).
3.
Gott, J. R. Closed timelike curves produced by pairs of moving cosmic
strings: exact solutions. Phys. Rev. Lett. 66, 1126–1129 (1991).
4.
Novikov, I. Evolution of the Universe Cambridge University Press:
Cambridge, England, (1983).
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1ιή22
31ή12ή2ί1ι
5.
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Hawing, S. W. Chronology protection conjecture. Phys. Rev. D 46, 603–611
(1992).
6.
Deutsch, D. Quantum mechanics near closed timelike lines. Phys. Rev. D 44,
3197–3217 (1991).
7.
Ralph, T. C. & Myers, C. R. Information flow of quantum states interacting
with closed timelike curves. Phys. Rev. A 82, 062330 (2010).
8.
Bacon, D. Quantum computational complexity in the presence of closed
timelike curves. Phys. Rev. A 70, 032309 (2004).
9.
Brun, T. A., Harrington, J. & Wilde, M. M. Localized closed timelike curves
can perfectly distinguish quantum states. Phys. Rev. Lett. 102, 210402
(2009).
10.
Ahn, D., Myers, C. R., Ralph, T. C. & Mann, R. B. Quantum state cloning in
the presence of a closed timelike curve. Phys. Rev. A 88, 022332 (2013).
11.
Brun, T. A., Wilde, M. M. & Winter, A. Quantum state cloning using
Deutschian closed timelike curves. Phys. Rev. Lett. 111, 190401 (2013).
12.
Pienaar, J. L., Ralph, T. C. & Myers, C. R. Open timelike curves violate
Heisenberg's uncertainty principle. Phys. Rev. Lett. 110, 060501 (2013).
13.
Bennett, C., Leung, D., Smith, G. & Smolin, J. A. Can closed timelike curves
or nonlinear quantum mechanics improve quantum state discrimination or
help solve hard problems? Phys. Rev. Lett. 103, 170502 (2009).
14.
Kłobus, W., Grudka, A. & Wójcik, A. Comment on ‘Information flow of
quantum states interacting with closed timelike curves’. Phys. Rev. A 84,
056301 (2011).
15.
Ralph, T. C. & Myers, C. R. Reply to ‘Comment on ‘Information flow of
quantum states interacting with closed timelike curves’’. Phys. Rev. A 84,
056302 (2011).
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1κή22
31ή12ή2ί1ι
16.
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Kitagawa, T. et al. Observation of topologically protected bound states in
photonic quantum walks. Nature Commun. 3, 882 (2012).
17.
Ma, X.-S. et al. Quantum simulation of the wavefunction to probe frustrated
Heisenberg spin systems. Nat. Phys. 7, 399–405 (2011).
18.
Simon, J. et al. Quantum simulation of antiferromagnetic spin chains in an
optical lattice. Nature 472, 307–312 (2011).
19.
Gerritsma, R. et al. Quantum simulation of the Dirac equation. Nature 463,
68–71 (2010).
20.
Casanova, J. et al. Quantum simulation of the Majorana Equation and
unphysical operations. Phys. Rev. X 1, 021018 (2011).
21.
Philbin, T. G. et al. Fiber-optical analog of the event horizon. Science 319,
1367–1370 (2008).
22.
Menicucci, N. C., Jay Olson, S. & Milburn, G. J. Simulating quantum effects
of cosmological expansion using a static ion trap. New J. Phys. 12, 095019
(2010).
23.
Lloyd, S. et al. Closed timelike curves via postselection: theory and
experimental test of consistency. Phys. Rev. Lett. 106, 040403 (2011).
24.
Lloyd, S., Maccone, L., Garcia-Patron, R., Giovannetti, V. & Shikano, Y.
Quantum mechanics of time travel through post-selected teleportation.
Phys. Rev. D 84, 025007 (2011).
25.
Brun, T. A. & Wilde, M. M. Perfect state distinguishability and
computational speedups with postselected closed timelike curves. Found.
Phys. 42, 341–361 (2011).
26.
Ralph, T. C., Langford, N., Bell, T. & White, A. G. Linear optical controlledNOT gate in the coincidence basis. Phys. Rev. A 65, 062324 (2002).
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
1λή22
31ή12ή2ί1ι
27.
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
Langford, N. K. et al. Demonstration of a simple entangling optical gate and
its use in bell-state analysis. Phys. Rev. Lett. 95, 210504 (2005).
28.
d'Espagnat, B. Conceptual Foundations of Quantum Mechanics 2nd edn
Addison Wesley (1976).
29.
Cavalcanti, E. G., Menicucci, N. C. & Pienaar, J. L. The preparation problem
in nonlinear extensions of quantum theory, Preprint at
http://arxiv.org/abs/1206.2725 (2012).
Acknowledgements
We thank Nathan Walk and Nicolas Menicucci for insightful discussions. We
acknowledge financial support from the ARC Centres of Excellence for
Engineered Quantum Systems (CE110001013) and Quantum Computation and
Communication Technology (CE110001027). A.G.W. and T.C.R. acknowledge
support from a UQ Vice-Chancellor's Senior Research Fellowship.
Author information
Affiliations
Centre for Engineered Quantum Systems, School of Mathematics and Physics,
University of Queensland, Brisbane, QLD 4072, Australia
Martin Ringbauer, Matthew A. Broome, Casey R. Myers & Andrew G. White
Centre for Quantum Computation and Communication Technology, School of
Mathematics and Physics, University of Queensland, Brisbane, QLD 4072, Australia
Martin Ringbauer, Matthew A. Broome, Andrew G. White & Timothy C. Ralph
Contributions
M.R., M.A.B., C.R.M. and T.C.R. developed the concepts, designed the experiment,
analysed the results and wrote the paper. M.R. performed the experiments and
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
2ίή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
analysed the data. T.C.R. and A.G.W. supervised the project and edited the
manuscript.
Competing interests
The authors declare no competing financial interests.
Corresponding author
Correspondence to Martin Ringbauer.
Supplementary information
PDF files
1.
Supplementary Information
υ π
α
α α 1-2, υ π
α
1-3 α
υ π
α
α απ
π
χό ια
π
α
α
, υ φ
α .Ε
α , πα α α
φ
α α
α
Επ ο νωνίες φύσης ISSN 2041-1723 (
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
α υ
υ
α α
απ υγ έαμ τθ
υ Ό
φ
α
υ
υ
α
υ
.
β)
21ή22
31ή12ή2ί1ι
ExperimentalΝsimulationΝofΝclosedΝtimelikeΝcurvesΝ|ΝσatureΝωommunications
© 2017 Macmillan Publishers Limited, ηΫλομ ου Springer Nature. Οζα α δεαδυηα α δα βλοτθ αδ.
httpsμήήwwwέnatureέcomήarticlesήncommsη14η
22ή22
WHAT IS A PHOTON?
Spontaneous emission: The need for quantum field theory
In these notes I would like to try and give an introduction to the quantum mechanical
theory of the photon. The treatment I give is in the spirit of a treatment you can find in
Dirac’s quantum mechanics monograph, The Principles of Quantum Mechanics. I believe
that Dirac was one of the first (if not the first) person to work out these ideas. Along
the way, we will be generalizing the way we use quantum mechanics in a non-trivial way.
More precisely, the way we model nature using the rules of quantum mechanics will change
significantly, although the rules themselves will not actually change. Let us begin by setting
the stage for this generalization.
Consider the well-known processes of emission and absorption of photons by atoms.
The processes of emission and absorption of photons by atoms cannot, ultimately, be
accommodated in the usual quantum mechanical models based on particle mechanics.
Instead, one must use a new class of models that go under the heading of quantum field
theory. The reasons for this necessity are relatively simple if one focuses on spontaneous
emission in atoms. This is where an atom in an excited state will spontaneously decay
to a lower energy state (and emit one or more photons). First, we all know that the
electron states we use to characterize atoms are the stationary states, which are energy
eigenstates. But stationary states have the property that all their observable features
are time independent. If an atomic electron occupying an atomic energy level were truly
in a stationary state there could never be any spontaneous emission since a stationary
state has no time dependent behavior. The only way out of this conundrum is to suppose
that atomic energy levels are not really stationary states once you take into account the
interaction of photons and electrons. But now consider emission of a photon by an atomic
electron. The initial state of the system has an electron. The final state of the system has
an electron and a photon. Now, in the usual quantum mechanical formalism for particles
the number of particles is always fixed. Indeed, the normalization of the wave function
for a particle (or for particles) can be viewed as saying that the particle (or particles) is
(or are) always somewhere. Evidently, such a state of affairs will not allow us to treat a
particle such as a photon that can appear and disappear. Moreover, it is possible to have
atomic transitions in which more than one photon appears/disappears. Clearly we will
not be able to describe such processes using the quantum mechanical models developed
thus far. If this isn’t surprising enough, I remind you that there exist situations in which
a photon may “transform” into an electron-positron pair and, conversely, in which an
electron-positron can turn into a photon. So even electrons are not immune from the
appearance/disappearance phenomenon.
1
Remarkably, it is possible to describe these multi-particle processes using the axioms
of quantum theory, provided these axioms are used in a clever enough way. This new
and improved use of quantum mechanics is usually called quantum field theory since it
can be viewed as an application of the basic axioms of quantum mechanics to continuous
systems (“field theories”) rather than mechanical systems. The picture that emerges is
that the building blocks of matter and its interactions are neither particles nor waves, but
a new kind of entity: a quantum field. Every type of elementary particle is described by
a quantum field (although the groupings by type depend upon the sophistication of the
model). There is an electron-positron field, a photon field, a neutrino field, and so forth.
In this way of doing things, particles are elementary excitations of the quantum field.
Quantum field theory (QFT) has lead to spectacular successes in describing the behavior of a wide variety of atomic and subatomic phenomena. The success is not just
qualitative; some of the most precise measurements known involve minute properties of
the spectra of atoms. These properties are predicted via quantum field theory and, so far,
the agreement with experiment is perfect.
We will only be able to give a very brief, very superficial, very crude introduction to
some of the basic ideas. Our goal will be to show how QFT is used to describe photons.
In a future discussion this could be the basis of an explanation of phenomena such as
spontaneous emission.
Harmonic Oscillators again
Our first goal will be to describe the photon without considering its interaction with
other (charged) particles. This is a quantum version of considering the EM field in the
absence of sources. Indeed, our strategy for describing photons will be to extract them
from a “quantization” of the source-free electromagnetic field. The key idea that allows
this point of view is that the EM field can, in the absence of sources, be viewed as an
infinite collection of coupled harmonic oscillators. We know how to describe harmonic
oscillators quantum mechanically, and we can try to carry this information over to the EM
field. First, let us quickly review the key properties of the harmonic oscillator.
Recall that the energy of a classical oscillator with displacement x(t) is given by
E=
m 2 1
ẋ + mω 2 x2 .
2
2
In the quantum description, we view the energy in terms of coordinate and momentum
operators, in which case the energy is known as the Hamiltonian:
H=
p2
1
+ mω 2 x2 .
2m 2
2
Here we view x and p as operators on a Hilbert space obeying the commutation relation
[x, p] = ih̄1,
with “1” being the identity operator. Let us recall the definition of the “ladder operators”:
r
r
p
p
mω
mω
†
(x + i
), a =
(x − i
),
a=
2h̄
mω
2h̄
mω
satisfying the commutation relations (exercise)
[a, a† ] = 1.
The Hamiltonian takes the form (exercise)
1
H = h̄ω(a† a + 1).
2
We can drop the the second term (with the “ 21 ”) if we want; it just defines the zero point
of energy. The stationary states are labeled by a non-negative integer n,
1
En = (n + )h̄ω.
2
H|ni = En |ni,
The ground state |0i satisfies
a|0i = 0.
Excited states are obtained via the identity (exercise)
a† |ni =
√
We also have
a|ni =
n + 1|n + 1i.
√
n|n − 1i.
It is easy to generalize this treatment to a system consisting of a number of uncoupled
harmonic oscillators with displacements xi , i = 1, 2, . . ., momenta pi , masses mi and
frequencies ωi . In particular, the Hamiltonian is (exercise)
!
X1
X †
p2i
1
1
2
2
H=
+ mω x
=
ai ai + 1 .
2 2mi 2 i i i
2
i
i
Even with a set of coupled oscillators, if the couplings are themselves harmonic, we can
pass to the normal mode coordinates and momenta. In this case the Hamiltonian again
takes the form given above. So, this description is quite general.
3
Fourier components of an EM field
To find a quantum description of the EM field we need some elementary results from
classical electromagnetic theory. To begin with, we introduce the EM potentials. Recall
that the homogeneous subset of the Maxwell equations
∇×E+
1 ∂B
=0
c ∂t
and
∇·B=0
are equivalent to the existence of a vector field, the vector potential A, and a scalar field,
the scalar potential φ, such that
E=−
1 ∂A
− ∇φ,
c ∂t
B = ∇ × A.
This is the general solution to the homogeneous subset of the Maxwell equations, so if we
choose to work with the electromagnetic potentials we have eliminated half of the Maxwell
equations.
The potentials are far from uniquely defined. If (φ, A) are a set of potentials for a
given EM field (E, B), then so are (exercise)
φ′ = φ −
1 ∂Λ
,
c ∂t
A′ = A + ∇Λ,
where Λ is any (well-behaved) function of space and time. This transformation between
two sets of potentials for the same EM field is called a gauge transformation, for historical reasons that we shall not go into. Physical quantities will be unchanged by gauge
transformations. The notion of gauge invariance, which just seems like a technical detail
in Maxwell theory, is actually pretty profound in physics. However, for our purposes, we
just note that the freedom to redefine potentials via a gauge transformation means that
we can try to make a convenient choice of potentials. Our choice will be always to put
the potentials in the radiation gauge. What this means is as follows. Any electromagnetic
field can be described by a set of potentials such that
φ = 0,
∇ · A = 0.
This should amuse you (at least a little). In electrostatics it is conventional to work in a
gauge in which A = 0 and then the static electric field is (minus) the gradient of the scalar
potential. This is certainly the most convenient way to analyze electrostatics, but one
could opt to use a time-dependent (and curl-free) vector potential if so-desired (exercise).
4
The Hamiltonian of the electromagnetic field
To use the harmonic oscillator paradigm to “quantize” the EM field, we first express
the total energy of the field in terms of the potentials:
"
#
2
Z
Z
1
1
1
∂A
d3 x (E 2 + B 2 ) =
d3 x 2
+ (∇ × A)2 .
H=
8π all space
8π all space
∂t
c
You can think of the first integral in the sum as the kinetic energy of the field and the
second integral as the potential energy. This is something more than an analogy. It is
possible to think of the electromagnetic field as a Hamiltonian dynamical system with
the vector potential playing the role of (an infinite number of) generalized coordinate(s)
and the electric field as its “canonically conjugate momentum”. In this interpretation the
function H above is the Hamiltonian (expressed in terms of position and velocity). The
behavior of any such dynamical system is determined by the Hamiltonian expressed as a
function of canonical coordinates and momenta, therefore we focus all our attention on H.
Next we make a Fourier decomposition of A:
Z
1
d3 k Ak (t)eik·x .
A(x, t) =
3/2
(2π)
Note that (exercises)
(Ak )∗ = A−k ,
∇ · A = 0 ⇐⇒ k · Ak = 0,
Z
1
∂A
d3 k Ȧk (t)eik·x ,
=
3/2
∂t
(2π)
Z
i
d3 k [k × Ak (t)]eik·x .
∇×A=
3/2
(2π)
Insert the Fourier expansion of the vector potential into the Hamiltonian. We get
(exercise):
Z
n1
o
H = d3 k 2 |Ȧk |2 + |k × Ak |2
c
Using (1) a vector identity for the dot product of a pair of cross products, and (2) the
radiation gauge condition, we get:
Z
o
n1
H = d3 k 2 |Ȧk |2 + k 2 |Ak |2 .
c
Let’s compare this with the harmonic oscillator Hamiltonian. Think of the integrand
as the energy of a single oscillator and the integral as a sum. Then we have (exercise)
ω −→ ω(k) = kc,
5
and
1
m −→ 2 .
c
The frequency correspondence is perfectly reasonable; it describes the frequency-wave number dispersion relation of an EM wave. The mass analogy is okay mathematically, but
shouldn’t be taken too literally in a physical sense; there is no particularly meaningful way
to ascribe a rest mass to the electromagnetic field.
Further confirmation of our interpretation of things comes by considering the remaining
Maxwell equations,*
1 ∂E
∇×B−
=0
c ∂t
and
∇ · E = 0,
which imply that each component of A satisfies the wave equation with propagation velocity c (and we still have the side condition ∇ · A = 0). This means that (exercise)
Ak (t) = ck eiω(k)t + c∗−k e−iω(k)t ,
for some constants ck . Thus for each value of k the vector potential behaves like 3 harmonic
oscillators with frequency kc.
Thus you can think of the vector potential as a sort of generalized coordinate, and the
electric field as the canonically conjugate momentum. The Hamiltonian is then that of
a (continuous) collection of harmonic oscillators labeled by the wave vector of a Fourier
decomposition. The fact that there is a continuous family of coordinates and momenta
(one for each spatial point, or one for each wave vector) leads to the statement that the
EM field has “an infinite number of degrees of freedom”. This is the principal feature
distinguishing quantum field theory from quantum mechanics, and it is what allows one
to describe processes in which particles are created and destroyed.
Now let us massage our formula for the Hamiltonian into an even nicer form. We do
this by taking account of the properties of the Fourier components of the vector potential.
We have seen that the Fourier component of the vector potential with wave vector k is
orthogonal to k, and that it satisfies a complex conjugation relation. We take both of these
conditions into account (and introduce a convenient normalization) via the definition
Ak =
2
X
σ=1
r
h̄c
(a ǫ
+ a∗−k,σ ǫ−k,σ ),
k k,σ k,σ
* Recall that we consider the electromagnetic field in regions of space in which there are no
charges or currents.
6
where σ labels the polarization and ǫk,σ are two orthonormal vectors orthogonal to k. The
variables ak,σ carry the amplitude information about each polarization, and the polarization direction is determined by the choice of ǫk,σ . Note that we have used h̄ to define the
new variables. From a purely classical EM point of view this is a bit strange, but it is
convenient for the quantum treatment we will give. For now, just think of the use of h̄ as a
convenient way of forming the amplitudes ak,σ , which are dimensionless (exercise). Using
this form of Ak (t) it follows that, for solutions of the Maxwell equations,
2
X
√
d
(ak,σ ǫk,σ − a∗−k,σ ǫk,σ ).
Ak = i h̄kc
dt
σ=1
Using this result we have (exercise)
Z
X
3
∗
H= d k
h̄ω(k) ak,σ ak,σ .
σ
Hopefully, this very simple form for the energy justifies to you all the effort that went
into deriving it. Up to a choice of zero point of energy, this is clearly a classical version of
the Hamiltonian for a collection of oscillators labeled by k and σ. A similar computation
with the Lagrangian for the Maxwell field,
Z
1
L=
d3 x(E 2 − B 2 ),
8π all space
leads to a sum (really, integral) of harmonic oscillator Lagrangians. Thus, mathematically
at least, the electromagnetic field (in the radiation gauge) is a continuous collection of
harmonic oscillators.
The quantization of the EM field
The classical Hamiltonian for the electromagnetic field can be expressed as a continuous
superposition over harmonic oscillator Hamiltonians:
Z
X
3
∗
H= d k
h̄ω(k) ak,σ ak,σ .
σ
We thus view the quantum EM field as an infinite set of quantum oscillators. The oscillators’ degrees of freedom are labeled by the wave vector k and the polarization σ. We view
†
the ladder operators for each degree of freedom as ak,σ and ak,σ . In the context of quantum field theory, we call these operators annihilation and creation operators, respectively.
We will see why in a moment. The quantum Hamiltonian is built from the creation and
annihilation operators via
Z
†
X
H = d3 k
h̄ω(k) ak,σ ak,σ .
σ
7
This is clearly just the sum of energies for each individual oscillator (with the “zero point
energy” dropped).
Incidentally, it is not too hard to compute the total momentum P of the electromagnetic field. It is obtained from the integral of the Poynting vector. (This means that the
Cartesian components of the total momentum are integrals of the corresponding components of the Poynting vector field. In terms of the creation and annihilation operators we
get (exercise)
Z
Z
†
X
c
3
d x E × B = d3 k
h̄k ak,σ ak,σ .
P=
4π
σ
This is the same as the total energy except that the energy of each mode, h̄ω(k) has been
replaced by the momentum of each mode h̄k.
Do you recall the usual lore of photons? You know, the lore that says a photon with a
definite energy and momentum will have E = h̄ω, and P = h̄k, where ω = kc. We see that
we are in a position to model a photon with definite energy and momentum as a quantum
normal mode of the EM field!
To make detailed sense of all this we should spell out the meaning of all these ladder
operators. The idea is to first use the harmonic oscillator point of view to understand the
definition of the operators. Then we can reinterpret the mathematical set-up in terms of
photon states.
The vacuum
To begin with, let us suppose that all the oscillators are in their ground state. This
state of the EM field, denoted |0i, is called the vacuum state. You can easily see why. This
state is an eigenstate of H and P with eigenvalue zero:
H|0i = 0 = P|0i.
This you can verify from the fact that all the annihilation (i.e., lowering) operators map
the ground state to the zero vector:
ak,σ |0i = 0.
Thus the vacuum of the EM field (in the absence of any other interactions, which we are not
modeling here) is a state in which the energy and momentum are known with probability
one. It can be shown that this state is the state of lowest energy and momentum.
By the way, have you ever encountered claims that, thanks to the uncertainty principle,
there is an “infinite reservoir of zero point energy in the vacuum”? Perhaps you have even
seen seemingly learned schemes to extract this energy for practical use. Now you are in
8
a position to be a little skeptical: the total energy-momentum of the vacuum is perfectly
well-defined – it vanishes! Where do these wacky claims come from? As with most popular
distortions of scientific results, there is a kernel of truth here. To uncover it, return to the
case of a single harmonic oscillator. The ground state energy is perfectly well defined, but
the energy and position and/or momentum operators are not compatible. This means that
if you know the energy with probability one, you cannot know the position or momentum
of the oscillator with probability one. Indeed, you may recall that in the ground state
of the oscillator the probability distributions for position and momentum are Gaussians
with zero average. The EM field has a similar behavior. Recall that the “position” is,
essentially, the vector potential (hence the magnetic field is a “function of position”) and
the “momentum” is, essentially, the electric field. In the EM vacuum state, the energy is
minimized and sharply defined, but the EM fields themselves “fluctuate”. More precisely,
the EM fields have a probability distribution with non-zero standard deviation; basically,
each mode (and polarization) is a described by a Gaussian probability distribution in the
vacuum state. (This fact is responsible for the “Casimir effect”.) Very roughly speaking,
the uncertainty principle means that, while the total energy is known with certainty, the
energy density is “uncertain”. I say “very roughly” since the notion of quantum energy
density of an EM field is rather touchy; there is no well-defined quantum version that
behaves much like the classical analog. Indeed, much of the zero point energy nonsense
that appears in print is based upon trying to use classical ideas to describe a feature of
the theory that is, well, not at all classical!
Photon states
With the vacuum state under control, we can now consider excited stationary states of
the quantum EM field, which we will interpret as states with one or more photons. Think
again about the EM field as a large collection of harmonic oscillators, labeled by wave
number and polarization. Suppose we put one of these oscillators in its first excited state,
†
say, by applying ak,σ to the vacuum for some fixed choice of k and σ. We denote this state
as
†
|1k,σ i = ak,σ |0i.
It can now be shown that the resulting state is an eigenvector of H and P:
H|1k,σ i = h̄ω(k)|1k,σ i,
P|1k,σ i = h̄k|1k,σ i.
This state has energy-momentum values defined with probability unity, which take the
form appropriate for a single photon. We interpret this state as a 1 photon state with
the indicated wave number, frequency, and polarization. Thus, the first excited stationary
9
state of the quantum electromagnetic field can be viewed as a single photon. In this
sense photons are “quanta of the electromagnetic field”. Superpositions of photon states
over momenta lead to photons that have probability distributions for their energy and
momenta. Thus photons need not have a definite momentum or energy any more than,
say, an electron must.
More generally, we can build up a Hilbert space of states by repeatedly applying to the
vacuum state the creation operators with various wave numbers and polarizations, taking
†
superpositions, etc. The result of each application of the creation operator ak,σ is to define
a state with one more photon of the indicated type. Each application of the annihilation
operator ak,σ results in a state with a photon of the indicated type removed. If the state
doesn’t have such a photon, (e.g., the vacuum state won’t) then the result is the zero
vector.
The states we are describing have a particle interpretation, but the theory is richer than
just a collection of particles since, e.g., one can superimpose the states described above
to get states which encode the field like properties of the EM field. Of course, such states
will not be stationary states. Also, recall our discussion of the “fluctuations” of E and
B in the vacuum. Charges respond to the electromagnetic field strengths (think: Lorentz
force law), and so the behavior of charged particles can be affected by EM phenomena,
even when no photons are present! This is the idea behind the “Lamb shift” found in the
spectra of atoms. And it is the key idea needed to explain spontaneous emission of photons
from atoms.
We see, then, how the “normal modes” of a field satisfying wave-type equations can be
“quantized”. The resulting theory admits particle-like properties in its stationary states.
To date, every known elementary particle has been successfully described by a quantum
field in much the way we described photons using a quantized electromagnetic field. Interactions between particles (better: between quantum fields) have also been described with
considerable success using the quantum field formalism. Indeed, it is reasonable to adopt
the point of view that, in our current best understanding of nature, quantum fields are the
stuff out of which everything is made!
10