Promoter diversity in multigene transformation

16
REVIEW Promoter diversity in multigene transformation Ariadna Peremarti Richard M. Twyman Sonia Go ´mez-Galera Shaista Naqvi Gemma Farre ´ Maite Sabalza Bruna Miralpeix Svetlana Dashevskaya Dawei Yuan Koreen Ramessar Paul Christou Changfu Zhu Ludovic Bassie Teresa Capell Received: 2 November 2009 / Accepted: 11 March 2010 / Published online: 31 March 2010 Ó Springer Science+Business Media B.V. 2010 Abstract Multigene transformation (MGT) is becoming routine in plant biotechnology as researchers seek to gen- erate more complex and ambitious phenotypes in trans- genic plants. Every nuclear transgene requires its own promoter, so when coordinated expression is required, the introduction of multiple genes leads inevitably to two opposing strategies: different promoters may be used for each transgene, or the same promoter may be used over and over again. In the former case, there may be a shortage of different promoters with matching activities, but repetitious promoter use may in some cases have a negative impact on transgene stability and expression. Using illustrative case studies, we discuss promoter deployment strategies in transgenic plants that increase the likelihood of successful and stable multiple transgene expression. Keywords Promoter Transgene Multigene transformation Transcriptional silencing Constitutive Spatiotemporal Inducible Introduction Genetic engineering has increased our fundamental under- standing of how plants function. By introducing new genes, we can study how plants grow, develop, and defend them- selves against pests, diseases and harsh environments, how photosynthesis is controlled, and the basis of primary and secondary metabolism (Slater et al. 2003). Genetic engineer- ing can also be applied to generate hardier crops, more nutri- tious food, and plants that produce chemical precursors, novel oils, industrial enzymes and pharmaceuticals (Farre et al. 2010; Zhu et al. 2007; Ma et al. 2003). The transferred genes play the most significant role in determining the phenotypes of transgenic plants but like stage hands working hard behind the scenes, the promoters that control transgene expression are essential components that often get overlooked. Promoters allow transgene expression to be regulated, restricted and fine- tuned, delivering more precise control over the manner in which phenotypes are expressed (Twyman 2003). The promoters used in plant biotechnology are tradition- ally divided into three categories—constitutive (active con- tinuously in most or all tissues), spatiotemporal (tissue- specific or stage-specific activity) and inducible (regulated by the application of an external chemical or physical signal; Potenza et al. 2004). However, all are based on similar core sequences usually including an initiator and TATA-box as well specific cis-acting motifs that bind to transcription fac- tors (Box 1). Promoter sequences are the same in all plant tissues whether the gene is expressed or not, so the activity of a promoter depends on the availability and activity of the transcription factors. Those binding to constitutive promoters are available and active all the time, whereas those binding to spatiotemporal and inducible promoters are themselves rationed and made available only in certain tissues or developmental stages, or in response to external signals. A. Peremarti S. Go ´mez-Galera S. Naqvi G. Farre ´ M. Sabalza B. Miralpeix S. Dashevskaya D. Yuan K. Ramessar P. Christou C. Zhu L. Bassie T. Capell Departament de Produccio ´ Vegetal i Cie `ncia Forestal, ETSEA, Universitat de Lleida, Av. Alcalde Rovira Roure 191, 25198 Lleida, Spain R. M. Twyman Department of Biological Sciences, University of Warwick, Coventry CV4 7AL, UK P. Christou (&) Institucio ´ Catalana de Recerca i Estudis Avanc ¸ats, Bellaterra, Spain e-mail: [email protected] 123 Plant Mol Biol (2010) 73:363–378 DOI 10.1007/s11103-010-9628-1

Transcript of Promoter diversity in multigene transformation

REVIEW

Promoter diversity in multigene transformation

Ariadna Peremarti • Richard M. Twyman • Sonia Gomez-Galera •

Shaista Naqvi • Gemma Farre • Maite Sabalza • Bruna Miralpeix •

Svetlana Dashevskaya • Dawei Yuan • Koreen Ramessar • Paul Christou •

Changfu Zhu • Ludovic Bassie • Teresa Capell

Received: 2 November 2009 / Accepted: 11 March 2010 / Published online: 31 March 2010

� Springer Science+Business Media B.V. 2010

Abstract Multigene transformation (MGT) is becoming

routine in plant biotechnology as researchers seek to gen-

erate more complex and ambitious phenotypes in trans-

genic plants. Every nuclear transgene requires its own

promoter, so when coordinated expression is required, the

introduction of multiple genes leads inevitably to two

opposing strategies: different promoters may be used for

each transgene, or the same promoter may be used over and

over again. In the former case, there may be a shortage of

different promoters with matching activities, but repetitious

promoter use may in some cases have a negative impact on

transgene stability and expression. Using illustrative case

studies, we discuss promoter deployment strategies in

transgenic plants that increase the likelihood of successful

and stable multiple transgene expression.

Keywords Promoter � Transgene �Multigene transformation � Transcriptional silencing �Constitutive � Spatiotemporal � Inducible

Introduction

Genetic engineering has increased our fundamental under-

standing of how plants function. By introducing new genes,

we can study how plants grow, develop, and defend them-

selves against pests, diseases and harsh environments, how

photosynthesis is controlled, and the basis of primary and

secondary metabolism (Slater et al. 2003). Genetic engineer-

ing can also be applied to generate hardier crops, more nutri-

tious food, and plants that produce chemical precursors, novel

oils, industrial enzymes and pharmaceuticals (Farre et al.

2010; Zhu et al. 2007; Ma et al. 2003). The transferred genes

play the most significant role in determining the phenotypes of

transgenic plants but like stage hands working hard behind the

scenes, the promoters that control transgene expression are

essential components that often get overlooked. Promoters

allow transgene expression to be regulated, restricted and fine-

tuned, delivering more precise control over the manner in

which phenotypes are expressed (Twyman 2003).

The promoters used in plant biotechnology are tradition-

ally divided into three categories—constitutive (active con-

tinuously in most or all tissues), spatiotemporal (tissue-

specific or stage-specific activity) and inducible (regulated

by the application of an external chemical or physical signal;

Potenza et al. 2004). However, all are based on similar core

sequences usually including an initiator and TATA-box as

well specific cis-acting motifs that bind to transcription fac-

tors (Box 1). Promoter sequences are the same in all plant

tissues whether the gene is expressed or not, so the activity of

a promoter depends on the availability and activity of the

transcription factors. Those binding to constitutive promoters

are available and active all the time, whereas those binding

to spatiotemporal and inducible promoters are themselves

rationed and made available only in certain tissues or

developmental stages, or in response to external signals.

A. Peremarti � S. Gomez-Galera � S. Naqvi � G. Farre �M. Sabalza � B. Miralpeix � S. Dashevskaya � D. Yuan �K. Ramessar � P. Christou � C. Zhu � L. Bassie � T. Capell

Departament de Produccio Vegetal i Ciencia Forestal, ETSEA,

Universitat de Lleida, Av. Alcalde Rovira Roure 191, 25198

Lleida, Spain

R. M. Twyman

Department of Biological Sciences, University of Warwick,

Coventry CV4 7AL, UK

P. Christou (&)

Institucio Catalana de Recerca i Estudis Avancats, Bellaterra,

Spain

e-mail: [email protected]

123

Plant Mol Biol (2010) 73:363–378

DOI 10.1007/s11103-010-9628-1

In the early years of plant biotechnology, most trans-

genic plants contained two transgenes—one selectable

marker under the control of a constitutive promoter to

facilitate the selective propagation of transformed cells,

and a ‘primary transgene’ which could be under the control

of any sort of promoter and was intended to alter the plant’s

phenotype in a specific manner. MGT is now being

embraced as an approach to generate plants with more

ambitious phenotypes (Naqvi et al. 2009a). MGT allows

goals that were once impossible to be achieved—the

import of complex metabolic pathways, the expression of

entire protein complexes, the development of transgenic

crops simultaneously engineered to produce a spectrum of

added-value compounds (Zhu et al. 2007; Naqvi et al.

2009a). But each additional transgene requires its own

promoter, making it necessary to find different promoters

that achieve the same expression profile or use one pro-

moter multiple times in the same transgenic plant. In the

first case there may be a shortage of available promoters

with the desired expression profile, whereas in the second

case the deliberate introduction of repetitive sequences into

a transgenic locus has in some cases been linked with

undesirable negative effects on transgene expression and

stability (Mette et al. 1999, 2000; Mourrain et al. 2007).

Promoter diversity

The challenge of multiple coordinated transgene expres-

sion can be addressed using a promoter diversity approach,

where different promoters are used to drive different

transgenes with the same expression profile. This is

achievable with most expression strategies when the

number of transgene is small, but becomes increasingly

difficult as the number of transgenes increases due to the

lack of available promoters with suitable expression pro-

files. Table 1 provides examples where MGT has been

carried out with transgenes driven in a coordinated manner

by different promoters.

Constitutive promoters

Constitutive promoters show the most diversity because

there are two major sources—plant viruses and plant

housekeeping genes. Plant viruses have small genes that

are easy to define genetically, and small genomes that are

easy to manipulate in vitro, so many of the earliest con-

stitutive promoters were derived from plant viruses and are

still widely used today. The most prevalent of these is the

Cauliflower mosaic virus 35S promoter, which controls the

synthesis of the 35S major transcript (Odell et al. 1985;

Kay et al. 1987). The full-sized regulatory region for the

CaMV 35S major transcript is just under 3 kbp in length

(Odell et al. 1985). However, shorter cloned fragments are

active, and the typical CaMV 35S promoter used in plant

expression vectors is a 352-bp fragment spanning nucleo-

tides -343 to ?9 (Fang et al. 1989). The core promoter

containing the TATA box extends to position -46 and

confers basal activity. Full activity requires upstream pro-

moter elements, some of which are tissue-specific, which

Box 1 Plant promoters, the basics

The core promoters of protein-encoding genes in plants are similar to those found in animals. They often contain a TATA box positioned at

approximately position -25 relative to the transcriptional initiation site, with the consensus sequence TATAWAW. They may also possess

an initiator element, which is found immediately adjacent to the transcriptional start site and has the consensus sequence YYCARR. These

two elements help to position the general transcription factor TFIID, which contains the TATA-binding protein (TBP) and various TBP-

associated factors (TAFs). However, some promoters lack both a TATA box and an initiator, and the genes under their control tend to have

multiple transcriptional initiation sites reflecting an inability to position TFIID precisely. Other components of the transcriptional initiation

complex then assemble: TFIIA, TFIIB, RNA polymerase II (recruited to the complex by TFIIF), TFIIE and TFIIH. These proteins have been

isolated and functionally characterized in a number of plants, and appear closely related to their animal counterparts (Twyman 2003).

Adjacent to the core promoter, further motifs are clustered in what is known as the upstream promoter region, which extends 100–200 bp on

the 50 side of the transcriptional start site. These motifs bind transcription factors that interact with the transcriptional initiation complex and

facilitate its assembly, improve its stability or increase the efficiency of promoter escape (movement away from the promoter, into the gene

itself) once the transcriptional machinery sets off. Some of these transcriptional activators are constitutive while others may be cell-specific,

stage-specific or inducible by external signals. Therefore, although the DNA sequence of the promoter is the same in each plant cell, the

availability of the transcription factors that bind to individual motifs varies considerably, and this provides the main mechanism for

transcriptional regulation. As well as the upstream promoter, there are more distant cis-acting elements known as enhancers that fulfill a

similar function. In virus genomes, enhancers are often found quite near the gene and promoter, allowing the isolation of all regulatory

elements on a relatively short DNA fragment. In plant genomes, enhancers may be located many kbp away, and may be found upstream,

downstream or even within the introns of the genes they control. Because of their distant location relative to the core promoter, interactions

with the transcriptional initiation complex are achieved by looping out the intervening DNA. Other proteins that bind to the upstream

promoter disrupt the transcriptional initiation complex, either directly or by blocking the activity of transcriptional activators. These

transcriptional repressors also bind at more distant sites known as silencers. Transcriptional activators and repressors function in various

ways to influence the stability of the initiation complex. Some are DNA binding proteins while others interact at the protein level. Some may

form direct contacts with the initiation complex, or indirect contacts through a bridging complex called mediator, while others modify

chromatin structure or introduce bends or kinks into the promoter DNA to facilitate or disrupt other interactions. Some regulatory proteins

can act as either activators or repressors depending on promoter context and the other proteins that are present.

364 Plant Mol Biol (2010) 73:363–378

123

Ta

ble

1E

xam

ple

so

fm

ult

igen

etr

ansf

orm

atio

n(M

GT

)w

ith

dif

fere

nt

pro

mo

ters

Nu

mb

er

of

gen

esa

Pro

mo

ters

So

urc

eA

ctiv

ity

Gen

e(a

nd

pro

du

ct)

Pat

hw

ay/f

un

ctio

nT

arg

etp

lan

tR

efer

ence

2A

9A

rab

ido

psi

sth

ali

an

aT

apet

um

-sp

ecifi

cB

ars

tar

Po

llen

dev

elo

pm

ent

Ind

ian

oil

seed

mu

star

d

(Bra

ssic

aju

nce

a)

Bis

ht

etal

.

(20

04)

TA

9T

ob

acco

Tap

etu

m-s

pec

ific

Ba

rsta

r

3P

DC

3W

(DC

3

pro

mo

ter

wit

hW

-en

han

cer

seq

uen

cefr

om

tob

acco

(Nic

oti

an

ata

ba

cum

))

Da

ucu

sca

rota

See

dsp

ecifi

cH

PP

D(4

-hy

dro

xy

ph

eny

lpy

ruv

ate

dio

xy

gen

ase)

Vit

amin

EC

ano

laR

acla

ruet

al.

(20

06)

Na

pin

Can

ola

(Bra

ssic

an

ap

us)

See

dsp

ecifi

cH

PT

(ho

mo

gen

tisa

tep

hy

tylt

ran

sfer

ase)

TC

(to

cop

her

ol

cycl

ase)

3P

7S

aS

oy

bea

nS

eed

-sp

ecifi

cA

t-H

PP

D(4

-hy

dro

xy

ph

eny

lpy

ruv

ate

dio

xy

gen

ase)

Vit

amin

EC

ano

laV

alen

tin

and

Mit

sky

(20

02)

P7

Sa

So

yb

ean

See

d-s

pec

ific

Eh-T

YR

A(c

ho

rism

ate

mu

tase

/pre

ph

enat

e

deh

yd

rog

enas

e)

PA

rc-5

Co

mm

on

bea

nS

eed

-sp

ecifi

cS

yn-V

TE

2(S

ynec

ho

cyst

issp

.P

CC

68

03

ho

mo

gen

tisa

tep

hy

tylt

ran

sfer

ase)

3U

sp(u

nk

no

wn

seed

pro

tein

pro

mo

ter)

Bro

adb

ean

(Vic

iafa

ba)

See

d-s

pec

ific

PtD

6(D

6-d

esat

ura

se)

PU

FA

syn

thes

isL

inse

ed(L

inu

mu

sita

tiss

imu

m)

Ab

bad

iet

al.

(20

04)

LeB

4(l

egu

min

)B

road

bea

nS

eed

-sp

ecifi

cP

SE

(D6

-elo

ng

ase)

Dc3

(hel

ian

thin

in)

D.

caro

taS

eed

-sp

ecifi

cP

tD5

(D5

-des

atu

rase

)

4P

7sa

(see

dst

ora

ge

pro

tein

pro

mo

ter)

So

yb

ean

(Gly

cin

em

ax)

See

d-s

pec

ific

At-

HP

PD

(p-h

yd

rox

yp

hen

ylp

yru

vat

e

dio

xy

gen

ase)

Vit

amin

ES

oy

bea

nK

aru

nan

and

aa

etal

.(2

00

5)

P7

saS

oy

bea

nS

eed

-sp

ecifi

cE

h-T

YR

A(b

ifu

nct

ion

alch

ori

smat

e

mu

tase

,p

rep

hen

ate

deh

yd

rog

enas

e)

PA

rc-5

(arc

elin

-5

pro

mo

ter)

Co

mm

on

bea

n

(Ph

ase

olu

svu

lga

ris)

See

d-s

pec

ific

At-

VT

E2

(ho

mo

gen

tisa

te

ph

yty

ltra

nsf

eras

e)

P7

Sa

So

yb

ean

See

d-s

pec

ific

At-

GG

H(g

eran

ylg

eran

yld

iph

osp

hat

e

hy

dra

tase

)

4O

leo

sin

(ole

osi

n

pro

mo

ter)

A.

tha

lia

na

See

d-s

pec

ific

At-

GG

H(g

eran

ylg

eran

yld

iph

osp

hat

e

hy

dra

tase

)

Vit

amin

ES

oy

bea

nK

aru

nan

and

aa

etal

.(2

00

5)

PF

AE

(fat

tyac

id

elo

ng

ase)

A.

tha

lia

na

See

d-s

pec

ific

At-

VT

E2

(ho

mo

gen

tisa

te

ph

yty

ltra

nsf

eras

e)

P7

Sa

So

yb

ean

See

d-s

pec

ific

At-

HP

PD

(4-h

yd

rox

yp

hen

ylp

yru

vat

e

dio

xy

gen

ase)

P7

Sa

So

yb

ean

See

d-s

pec

ific

Eh-T

YR

A(c

ho

rism

ate

mu

tase

/pre

ph

enat

e

deh

yd

rog

enas

e)

Plant Mol Biol (2010) 73:363–378 365

123

Ta

ble

1co

nti

nu

ed

Nu

mb

er

of

gen

esa

Pro

mo

ters

So

urc

eA

ctiv

ity

Gen

e(a

nd

pro

du

ct)

Pat

hw

ay/f

un

ctio

nT

arg

etp

lan

tR

efer

ence

4L

MW

glu

ten

inW

hea

tE

nd

osp

erm

Zm

psy

1(p

hy

toen

esy

nth

ase

1)

Car

ote

no

id,

asco

rbat

e

and

fola

te

Co

rnN

aqv

iet

al.

(20

09

b)

D-h

ord

ein

Bar

ley

En

do

sper

mP

acr

tI(p

hy

toen

ed

esat

ura

se)

Osd

ha

r(d

ehy

dro

asco

rbat

ere

du

ctas

e)

Ecf

olE

(GT

Pcy

clo

hy

dro

lase

I)

5L

MW

G(l

ow

mo

lecu

lar

wei

gh

tg

lute

nin

)

Wh

eat

(Tri

ticu

ma

esti

vum

)

En

do

sper

mp

sy1

(ph

yto

ene

syn

thas

e1

)C

aro

ten

oid

Co

rnZ

hu

etal

.(2

00

8)

Ho

rP(h

ord

ein

)B

arle

y(H

ord

eum

vulg

are

)

En

do

sper

mcr

tI(p

hy

toen

ed

esat

ura

se)

RP

5(p

rola

min

)R

ice

(Ory

zasa

tiva

)E

nd

osp

erm

lycb

(ly

cop

ene

b-cy

clas

e)

Gt1

(glu

teli

n-1

)R

ice

En

do

sper

mb

ch(b

-car

ote

ne

hy

dro

xy

lase

)

GZ

63

(c-z

ein

)C

orn

(Zea

ma

ys)

En

do

sper

mcr

tW(b

-car

ote

ne

ket

ola

se)

7P

FA

E(f

atty

acid

elo

ng

ase

1)

A.

tha

lia

na

See

d-s

pec

ific

crtW

(b-C

aro

ten

ek

eto

lase

)

crtB

(ph

yto

ene

syn

thas

e)

Car

ote

no

idC

ano

laF

uji

saw

aet

al.

(20

09)

Na

pin

Can

ola

See

d-s

pec

ific

idi

(iso

pen

ten

yl

py

rop

ho

sph

ate

iso

mer

ase)

crtZ

(b-c

aro

ten

eh

yd

rox

yla

se)

CaM

V3

5S

Co

nst

itu

tiv

ecr

tE(g

eran

ylg

eran

yl

py

rop

ho

sph

ate

syn

thas

e)

crtI

(ph

yto

ene

des

atu

rase

)

crtY

(ly

cop

ene

b-c

ycl

ase)

aE

xcl

ud

ing

the

sele

ctab

lem

ark

erg

ene

366 Plant Mol Biol (2010) 73:363–378

123

means that constitutive expression is achieved through the

additive effects of multiple tissue-specific motifs (Lam

et al. 1989). The region between nucleotides -90 and

-208 is an enhancer, which can increase promoter activity

in a linear fashion when up to four copies are present (Kay

et al. 1987). The enhancer can function in either orientation

and at different positions relative to the gene (Fang et al.

1989; Ohtsuki et al. 1998), and can also enhance the

activity of heterologous promoters to which it is attached

(Zheng et al. 2007).

Although widely used, the CaMV 35S promoter has

certain limitations such as its poor performance in mono-

cots, its suppression by feeding nematodes (Goddijn et al.

1993; Urwin et al. 1997) and the intellectual property

issues affecting its commercial deployment. For this rea-

son, alternative virus promoters with similar or improved

properties have been sought. Because the CaMV 35S

promoter has proven so successful, related caulimoviruses

have been the first port of call. Examples include promoters

from Figwort mosaic caulimovirus (FMV; Bhattacharyya

et al. 2002), Cassava vein mosaic virus (CsVMV; Verda-

guer et al. 1996), Cestrum yellow leaf curling virus

(CmYLCV; Stavolone et al. 2003), Mirabilis mosaic virus

(MiMV; Dey and Maiti 1999) and Peanut chlorotic streak

virus (PClSV; Maiti and Shepherd 1998; Bhattacharyya

et al. 2003). The CmYLCV promoter sequence is available

from Syngenta Biotechnology, Inc. for limited research

purposes, and has been used to express glyoxalase I (gly I)

to achieve salt tolerance in tobacco (Veena et al. 1999) and

mung bean (Bhomkar et al. 2008).

Badnaviruses are closely related to the caulimoviruses,

and they contain a strong constitutive promoter that tran-

scribes the entire genome. The Commelina yellow mottle

virus (CoYMV) promoter is active in tobacco (Medberry

et al. 1992), but more detailed studies have been carried out

in monocots where its activity is broad albeit not universal

(Torbert et al. 1998). The Sugarcane bacilliform virus

(ScBV) promoter is active in monocots (banana, corn,

millet and sorghum) and dicots (tobacco, sunflower, canola

and Nicotiana benthamiana) and has been shown to drive

high level gusA expression in transgenic banana and

tobacco plants (Schenk et al. 2001). A subcloned fragment

of the promoter was also shown to drive gusA expression in

sugarcane to at least the same level as achieved with the

corn Ubi-1 promoter (Braithwaite et al. 2004).

Subterranean clover stunt virus (SCSV) is a member of

the Circoviridae and has eight circular genome segments

each containing a promoter. Initial studies showed that the

promoters from segments four and seven were comparable

in strength to the CaMV 35S promoter in cotton, potato and

tobacco, and further analysis showed they were also suit-

able for expression in monocots, giving rise to the pPLEX

series of expression vectors (Schunmann et al. 2003a, b).

Plant housekeeping genes are another important source

of constitutive promoters, since housekeeping genes

encode proteins that are required by all cells for basic

functions such as core metabolism and maintenance of cell

structure and integrity. One major class of housekeeping

genes sees to the synthesis of cytoskeletal components, and

includes large families of genes encoding actins and tub-

ulins. The rice actin1 promoter drives strong transgene

expression in rice protoplasts transiently expressing gusA

(McElroy et al. 1990) and in most tissues of transgenic rice

plants (Zhang et al. 1991). Interestingly, McElroy and

colleagues showed that transient expression was abolished

in the absence of the first intron of the actin1 gene

(McElroy et al. 1990) and that the inclusion of this intron in

a chimeric CaMV 35S promoter enhanced its activity in

transgenic rice and corn plants by 40-fold (McElroy et al.

1991). In contrast, the first intron of the recently charac-

terized rice actin2 gene contains a negative regulatory

element whose removal is required for high-level promoter

activity; 2.6 kbp of upstream sequence contains all the 50

regulatory elements necessary for high-level constitutive

gusA expression in transgenic rice (He et al. 2009a). Other

actin promoters that have been used for constitutive or

near-constitutive expression include Arabidopsis ACT2

(An et al. 1996) and banana ACT1 (Hermann et al. 2001).

The ubiquitins are another highly conserved family of

housekeeping genes, some constitutively expressed (Ka-

walleck et al. 1993) others responsive to stress (Christensen

et al. 1992; Cornejo et al. 1993; Christensen and Quail

1996). For example, the Arabidopsis UBQ1 and UBQ6

promoters are active in all tobacco tissues, albeit at dif-

ferent levels (Callis et al. 1990), and the corn Ubi-1 pro-

moter is more than 10 times stronger than the CaMV35S

promoter in corn protoplasts when combined with the first

intron (Norris et al. 1993). Significant improvements in

activity have also been observed with three Arabidopsis

polyubiquitin promoters (UBQ3, UBQ10 and UBQ11) and

the potato ubi7 promoter (Garbarino et al. 1995).

Other housekeeping genes are involved in protein syn-

thesis and core metabolism. One example is NeIF4A-10, a

member of the tobacco eIF4A gene family—even a small

region of the promoter (to position -188) drives consti-

tutive expression in transgenic tobacco plants (Mandel

et al. 1995). Further characterization of the promoter

revealed a strong enhancer in the region -151 to -73 and

showed that an intron is not required to boost promoter

activity (Tiana et al. 2005). Other examples are the pro-

moters for genes encoding acetolactate synthase (ALS),

which catalyses the first step in the biosynthesis of the

branched chain amino acids leucine, isoleucine and valine,

and ACC synthase, which plays an important role in eth-

ylene synthesis. The Brassica ALS3 promoter is constitu-

tive and comparable in strength to the CaMV 35S promoter

Plant Mol Biol (2010) 73:363–378 367

123

(Baszczynski et al. 1997). VR-ACS1 is an auxin-inducible

ACC synthase gene from mung bean, whose promoter is

constitutive and 4–6 times more active than CaMV 35S in

Arabidopsis and tobacco (Cazzonelli et al. 2005). It is not

clear why the promoter is constitutive in a heterologous

background when it is tightly regulated in mung bean, but

this probably reflects the absence of a critical transcrip-

tional repressor in the transgenic plants.

With such a spectrum of diverse constitutive promoters

available it is perhaps surprising that there are so few

examples where different promoters have been incorpo-

rated into the same transgenic plant to drive multiple

transgenes (Table 1). In most of these examples, the CaMV

35S promoter has been used to regulate one transgene, with

the other(s) driven by a different viral or plant constitutive

promoter. However, there are cases where multiple pro-

moters other than the CaMV 35S promoter have been used

particularly in commercial varieties where the IP con-

straints surrounding the CaMV 35S promoter would make

commercial development more challenging.

Spatiotemporal promoters

Many different plant promoters have been described that

restrict expression to particular cells, tissues, organs or

developmental stages. Within this collection the number of

different promoters that can be used to drive coordinated

transgene expression (multiple transgenes with the same

expression profile) is somewhat more limited, and case

studies in which the promoter diversity strategy has been

applied are therefore fewer in number (Table 1).

Arguably the most abundant spatiotemporal promoters

are those that restrict transgene expression to seeds, and for

many reasons the seeds are often a favored target for

transgene expression, particularly if the goal of an experi-

ment is to force the accumulation of a heterologous product

that might interfere with vegetative growth at high con-

centrations or to improve the nutritional quality of seeds

used as staple foods. Many promoters have been identified

that target genes specifically to the seed, or to a particular

region of the seed. Storage proteins such as corn zein

(Schernthaner et al. 1988), rice glutelin (Leisy et al. 1989;

Takaiwa et al. 1991; Zheng et al. 1993), barley hordein

(Marris et al. 1988), rice prolamin (Qu and Takaiwa, 2004)

and wheat glutenin (Colot et al. 1987) have been rich

sources of seed-specific promoters, predominantly directing

expression to the endosperm (Wobus et al. 1995). Addi-

tional promoters have been shown to direct gene expression

to the embryo and aleurone (Opsahl-Sorteberg et al. 2004;

Qu and Takaiwa 2004; Furtado and Henry 2005).

The spatial and temporal activities of seed-specific

promoters reflect their peculiar combination of cis-acting

regulatory motifs (Onodera et al. 2001). The endosperm-

specific bZIP transcription factors O2 (Opaque-2; corn) and

SPA (storage protein activator; wheat) have been shown to

activate prolamin genes by binding to the GCN4-like motif

in the bifactorial endosperm box (Albani et al. 1997;

Schmidt et al. 1992). The GCN4 motif, a cis-acting ele-

ment that is highly conserved in the promoters of cereal

seed storage protein genes, plays a central role in con-

trolling endosperm-specific expression (Wu et al. 1998).

The barley endosperm-specific bZIP transcriptional acti-

vators BLZ1 and BLZ2 activate storage protein genes by

interacting with the GCN4 motif in barley B1-hordein

promoters (Vicente-Carbajosa et al. 1998; Onate et al.

1999), and the rice bZIP transcription factor, RISBZ1

works in a similar manner (Onodera et al. 2001). It has

been demonstrated that GCN4 and AACA motifs are

conserved in all glutelin gene promoters isolated thus far

(Takaiwa et al. 1996), and the GCN4 motif is also observed

in the PPK, SBE1 and 10, 13 (PG5a) and 16 kDa prolamin

promoters. This motif acts as a key element conferring

aleurone and subaleurone-specific expression (Wu et al.

1998; Qu and Takaiwa 2004).

Anther-specific promoters are also very useful because

they can be used to control male fertility, an important trait

in crop breeding. Numerous anther-specific and pollen-

specific promoters have been identified in a variety of

plants, including the TA29 promoter from tobacco (Kol-

tunow et al. 1990), the A9 promoter from Arabidopsis (Paul

et al. 1992) and the RA8 promoter from rice (Jeon et al.

1999). A cis-acting element found between 285 and

2,207 bp upstream of the TA29 gene confers strong, tape-

tum-specific expression and appears to work in heterolo-

gous settings, as it has been used to create male-sterile

tobacco and rapeseed plants through the expression of

Bacillus amyloliquefaciens barnase, a potently cytotoxic

ribonuclease (Mariani et al. 1990). Fertility can be restored

in F1 plants if female barnase transgenic parents are cros-

sed with males expressing the barnase inhibitor barstar

(Mariani et al. 1992), but the frequency of full restoration

can be low due to poor control over transgene expression.

Bisht et al. (2004, 2007) addressed this by enhancing and

augmenting the expression of barstar using a combination

of gene optimization, promoter engineering and MGT. The

wild-type and synthetic barstar genes were expressed using

either the TA29 and A9 promoters separately, or a chimeric

promoter combining elements from each. Although the

chimeric promoter was more active than either of the

individual promoters, the expression of different versions

of barstar using the two different tapetum-specific pro-

moters achieved the greatest frequency of restoration to

fertility.

Many case studies have been published in which

multiple transgenes are expressed in specific tissues by

combining the use of spatiotemporally-regulated and

368 Plant Mol Biol (2010) 73:363–378

123

constitutive promoters but there have been few examples of

studies in which different spatially-restricted promoters

have been used in the same plant. The barnase/barstar case

study described above is one example (Bisht et al. 2004,

2007). Another involved the use of five different promoters

to drive five transgenes in corn endosperm, i.e., corn

phytoene synthase 1, Pantoea ananatis phytoene desatur-

ase, Gentiana lutea lycopene b-cyclase, G. lutea b-caro-

tene hydroxylase, and Paracoccus sp. b-carotene ketolase,

controlled by the low molecular weight wheat glutenin,

barley hordein, rice prolamin, rice glutelin-1 and corn

c-zein promoters, plus the selectable marker gene under

constitutive control (Zhu et al. 2008). This study showed

how combinatorial transformation with multiple genes

could be used to generate a library of plants with different

phenotypes representing carotenoid biosynthesis. The dif-

ferent promoters ensured that, in any combination, it would

be possible to isolate plants without multiple copies of the

same promoter thus reducing the likelihood of transcrip-

tional silencing.

Inducible promoters

Arguably the most useful promoters in plants are those

responsive to the changing environment. Many different

inducible promoters have been identified in plants and

these generally fall into three categories—(1) those

responsive to endogenous signals (plant hormones); (2)

those responsive to external physical stimuli (abiotic and

biotic stresses); and (3) those responsive to external

chemical stimuli. Such promoters provide immense scope

for the precise regulation of transgene expression through

external control, ranging from the precise control of

transgene activation/inactivation in experimental settings

to the ability to activate transgenes on an agricultural scale

by the application of chemical sprays. Like constitutive

promoters, there are two major sources for inducible pro-

moters—endogenous genes and heterologous systems.

Inducible promoters have been combined with both con-

stitutive and spatiotemporal promoters in transgenic plants

but are generally not combined with each other, i.e., where

multiple genes need to be induced together it is standard

practice to place them under the control of the same pro-

moter. The inducible promoters most commonly used for

MGT strategies include hormone responsive promoters,

heat shock promoters and those responding to pathogens.

The most widely-used hormone-responsive promoters

are those induced by auxins, gibberellins and abscisic acid,

although promoters responsive to heterologous hormones

(from insects and mammals) are also useful because they

do not activate endogenous pathways. For example, Mar-

tinez et al. (1999) developed a hybrid system consisting of

the tobacco budworm ecdysone receptor ligand-binding

domain fused to the mammalian glucocorticoid receptor

DNA-binding domain and the VP16 transactivation

domain. The receptor responds to tebufenozide (an insec-

ticide better known by its trade name CONFIRM). Simi-

larly, Padidam et al. (2003) have developed a system that is

based on the spruce budworm ecdysone receptor ligand-

binding domain, and responds to another common insec-

ticide, methoxyfenozide (INTREPID). Another system

based on the European corn borer ecdysone receptor also

responds to this insecticide (Unger et al. 2002).

The application of exogenous auxins induces certain

genes within minutes so auxin-inducible promoters are

suitable for rapid responses. The cis-acting auxin response

elements in these genes comprise the motif TGTCTC adja-

cent to or overlapping a constitutive motif, such that the

auxin-response element is occupied and the constitutive

element repressed when the auxin concentration is low, and

the opposite occurs when the concentration is higher. Syn-

thetic auxin response elements can be engineered to contain

multiple TGTCTC repeats, and two domains from the pea

PS-IAA4/5 promoter can induce transcription from the

CaMV 35S basal promoter following auxin application

(Ballas et al. 1995). Additional auxin response elements

include the as-1/ocs-like element (TGACG(T/C)AAG(C/

G)(G/A)(C/A) T(G/T)ACG(T/C)(A/C)(A/C); Ellis et al.

1993). Auxin-response elements are often found combined

with response elements to other hormones, allowing com-

plex control. For example, the promoter of the brassinolide-

inducible BLE3 gene in rice also contains auxin response

elements which allow dual regulation by brassinolides and

IAA (Yang et al. 2006). Gibberellins also induce gene

expression rapidly, but in this case signaling is achieved

when the presence of gibberellic acid (GA) induces the

degradation of DELLA proteins, which bind to silencer

elements in GA-inducible genes and interfere with the

binding of transcriptional activators to GA response ele-

ments (GARE). In the absence of DELLA proteins, the GA-

regulated MYB transcription factor (GAMYB) binds to

GAREs and activates transcription (Sun and Gubler 2004;

Woodger et al. 2003). GA-regulated gene expression has

been studied exhaustively in the barley a-amylase genes,

revealing the core of the GARE to be a TAACAAA-like box

which can be repressed by abscisic acid (ABA). Skriver et al.

(1991) reported that when the GARE is fused as a tandem

repeat of six units it can confer GA responsiveness to a

CaMV 35S minimal promoter. Other studies have identified

further elements that lie a conserved distance from the

GARE and, in different combinations, help to modify the

expression profile. These elements combined with the

GARE constitute what is known as a GAR complex, which

can include the pyrimidine box (C/TCTTTT), the TATC-

CAC box, the CAACTC box and the Box1/O2S-like ele-

ment. Promoters that are inducible by ABA usually contain

Plant Mol Biol (2010) 73:363–378 369

123

one or more ABA response elements (ABREs), which may

include a G-box/C-box as well as MYC and MYB binding

sites. There have been many reports of promoters whose

activity has been modified by the incorporation of such

elements. Sequence analysis of the cotton late embryogen-

esis-abundant (LEA) D113 gene promoter (Luo et al. 2008)

revealed the presence of ABREs, a drought-response ele-

ment, and MYB and MYC sites; a 158 bp fragment was

shown to be sufficient for ABA induced reporter gene

expression. Some promoters can also be negatively regu-

lated by ABA. The corn HyPRP gene is specifically

expressed in immature embryos and its expression is

inhibited when ABA levels increase during maturation. A

construct containing 2 kbp of the HyPRP promoter joined to

the gusA gene allowed the repression of GUS activity in corn

and tobacco following the application of ABA. Two ACGT

elements at positions -260 and -93 bp in the promoter are

likely to mediate this effect (Jose-Estanyol et al. 2005).

Many metabolic genes need to respond sensitively to the

levels of specific metabolites to facilitate feedback regu-

lation loops that tie gene expression to metabolic balance.

Therefore, many promoters respond to the presence of

certain chemicals, and these can be extremely useful to

regulate transgene expression in plants because unlike

hormones they do not have pleiotropic effects on growth

and development. Sugar responsive promoters contain a

variety of response elements and these can confer sugar-

sensitivity on heterologous genes. For example, elements

from sporamin and amylase promoters have been studied in

detail and the minimal a-amylase 3 promoter makes the

normally constitutive rice actin1 promoter sensitive to the

presence of sugar (Lu et al. 1998). SURE (SUcrose

REsponsive) elements are found in sugar-inducible patatin

promoters (Grierson et al. 1994; Zourelidou et al. 2002)

and in the promoter of the VvTH1 gene, which encodes a

hexose transporter (Atanassova et al. 2003).

Plant defense genes often have promoters that are

inducible by chemicals (elicitors) produced by pests and

pathogens. Different combinations of trans-acting factors

are responsible for the elicitor-mediated activation of

defense related genes and numerous cis-acting elements

have been identified in the corresponding promoters. They

can be found as single or multiple copies and in different

combinations, allowing some genes to respond to more

than one elicitor, and particular pathogens to elicit distinct,

functionally-related sets of genes. The cis-acting elements

are highly conserved among different plant species so they

are very useful for making transgenes pathogen-inducible

in a range of heterologous settings. Individual elicitor

response elements appear to work well in a different pro-

moter context allowing a certain degree of ‘plug-and-play’

between promoters. The number of copies of each element

also appears to have an effect on the strength of induction,

thus promoters with one or two copies are best suited to

mediate pathogen-specific expression (Rushton et al.

2002). Many plants also respond to the physical aspects of

pest activity (wounding) by expressing a set of protective

genes, many of which are also inducible by jasmonate or

ethylene. Wound-inducible genes include those encoding

proteinase inhibitor II (pinII; Godard et al. 2007; Keinon-

en-Mettala et al. 1998; Keil et al. 1989; Xu et al. 1993), the

Agrobacterium tumefaciens enzymes nopaline synthase

(An et al. 1990) and mannopine synthase (Guevara-Garcıa

et al. 1993; Ni et al. 1996), and peroxidases such as rice

R2329 (Sasaki et al. 2007). The use of wound-inducible

promoters in transgenic plants shows that wound signaling

pathways are functionally conserved across taxonomic

groups (Wu et al. 1999; Yevtushenko et al. 2004).

Elevated temperatures result in the expression of so-

called heat shock proteins (HSPs), which protect plants

from the effects of heat by helping to stabilize proteins and

other cellular components thus conferring thermotolerance.

HSP genes have heat-inducible promoters that are also

strongly conserved across species. For example, the soy-

bean hsp17.3B promoter was used to drive gusA expression

in the moss Physcomitrella patens resulting in no GUS

activity at 25�C but increasing GUS activity at higher

temperatures (Saidi et al. 2005). Heat shock elements

(HSEs) were identified in the tomato ftsH6 promoter, and

were recognized by another HSP acting as a transcription

factor (heat shock factor, HSF; Sun et al. 2006). The heat

shock response may be universal or tissue specific. The

Arabidopsis GolS1 promoter confers universal GUS

activity in transgenic Arabidopsis plants following a tem-

perature shift (Panikulangara et al. 2004), whereas the

barley hspl7 promoter confers heat-induced expression

only in the xylem bundles of the stem and petioles—this

expression profile is conserved in transgenic monocot and

dicot plants (Raho et al. 1996).

Repetitious promoters and transcriptional

gene silencing

The promoter diversity approach discussed above is

employed to avoid repetitious use of the same promoter,

which can in some cases promote transcriptional gene

silencing through promoter methylation and inactivation.

This effect is usually constitutive but may be tissue-specific

(Kloti et al. 2002). However, there have been plenty of

reports describing transgenic plants carrying multiple

transgenes under the control of the same promoter yet

showing strong and stable expression. For example,

although Zhu et al. (2008) used five different endosperm-

specific promoters in corn to achieve the high-level

expression of carotenogenic genes, Naqvi et al. (2009b)

370 Plant Mol Biol (2010) 73:363–378

123

achieved strong expression of four genes in the same sys-

tem using the barley D-hordein promoter to control each

transgene, with no adverse effects. The literature suggests

that repetitious use of the same promoter is often suc-

cessful, so what are the exact limitations of using the same

promoter to control more than one transgene?

To answer this question, it is necessary to look more

closely at the reasons for gene silencing caused by repetitive

promoter use. In plants, transcriptional gene silencing

resulting from sequence homology in promoter regions is

correlated with increased promoter methylation (Kooter

et al. 1999) and appears to be driven by the production of

double-stranded RNA (dsRNA) matching the promoter

sequence (Mette et al. 2000). This has been demonstrated

by deliberately expressing dsRNA corresponding to the nos

promoter in transgenic plants carrying a second transgene

driven by the nos promoter (Mette et al. 1999) and by

constructing transgenic plants in which the transgene locus

triggers both transcriptional and post-transcriptional

silencing simultaneously, by producing dsRNA corre-

sponding to the promoter and transcribed sequences of

different target genes (Mourrain et al. 2007). The compo-

nents of this RNA-driven DNA methylation system have

been systematically sought and investigated, and include

RNA-dependent RNA polymerase 2 (RDR2) and Dicer-like

3 (DCL3; Xie et al. 2004), Pol IV (Herr et al. 2005; Kanno

et al. 2005; Onodera et al. 2005), Pol IVb/Pol V (Pontier

et al. 2005; Wierzbicki et al. 2008; He et al. 2009b), Arg-

onaute 4 (AGO4; Zilberman et al. 2003), chromatin

remodeling protein DRD1 (Kanno et al. 2004), and de novo

DNA methyltransferase DRM2 (Cao and Jacobsen 2002).

In the absence of deliberately-created promoter dsRNA,

the transcriptional silencing seen in some transgenic plants

carrying multiple copies of the same promoter appears to

arise from dsRNA produced serependitiously. The organi-

zation of a transgenic locus is difficult to control, and it is

therefore a common occurrence in both Agrobacterium-

mediated transformation and direct DNA transfer that the

juxtaposition of transgenes or fragments thereof can result

in the creation of hairpin promoter structures at the DNA

level that are transcribed into aberrant dsRNA species.

Such arrangements can be obvious and easy to detect, but

even where gross rearrangements are absent it is possible

that undetected ‘micro-rearrangements’ are present in the

transgenic locus, as observed by Mehlo et al. (2000) when

investigating the structure of a transgenic locus in corn

generated by direct DNA transfer. The siRNAs that trigger

RNA-dependent DNA methylation are just 24 bp in length,

so it is conceivable that inverted repeats of \50 bp could

be sufficient for transgene silencing, and such structures

would be undetectable using the coarse analysis methods

typically employed to study transgenic plants, such as

Southern blot hybridization.

There are at least four mechanisms by which promoter-

specific hairpin RNAs could be generated in transgenic

plants carrying different transgenes under the control of the

same promoter. In many cases, the creation of hairpin RNA

structures may be a consequence of the relative position of

intact transgenes, which is itself a reflection of the mecha-

nism of transgene integration. The organization of integrated

T-DNA sequences differs among Agrobacterium strains, but

a common feature of nopaline-type derivatives such as C58

is the preferential integration of T-DNA as dimers with an

inverted repeat configuration, linked either at the left or right

borders (Jones et al. 1987; Jorgensen et al. 1987). Where

cotransformation is carried out with two T-DNAs containing

different genes, the different T-DNAs often integrate as

heterodimeric inverted repeats, preferentially around the

right border (De Block and Debrouwer 1991). If the same

promoter is used for both genes, this would favor the for-

mation of hairpin structures that could be transcribed from

the opposite strand. The structure of loci generated by direct

DNA transfer is more variable, but inverted repeat structures

involving promoter sequences are not uncommon, allowing

the same silencing mechanism to operate (Kohli et al. 2003).

The second mechanism reflects the frequent occurrence

of spontaneous transgene rearrangements (truncation,

fragmentation, recombination) which can be thought of as

‘collateral damage’ occurring to DNA during the transfor-

mation process due to the activities of nucleases and repair

complexes that facilitate the integration of exogenous DNA.

There have been few systematic studies of this phenomenon

in Agrobacterium-mediated transformation, but Afolabi

et al. (2004) and Zhu et al. (2006) found that nonintact

T-DNAs were present in [70% of transgenic rice lines, in

most cases reflecting loss of the mid to right border portion

of the T-DNA. Similarly, Rai et al. (2007) found that about

50% of rice plants transformed with a T-DNA containing

the phytoene synthase (psy) and phytoene desaturase (crtI)

genes showed evidence of rearrangements, and in the

majority of cases the rearrangements occurred in the crtI

expression cassette, which was adjacent to the right T-DNA

border. Rearrangements involving the left border are often

characterized by the insertion of variable-size regions of

filler DNA, which may be derived from the T-DNA

sequence or from plant genomic DNA, providing further

scope for the creation of hairpin structures (De Buck et al.

1999; Kumar and Fladung 2000, 2002).

A third mechanism for silencing triggered by repetitive

promoter usage reflects the structure of the promoter itself.

Rearrangements affecting the promoter of a transgene may

be more common than expected due to the presence of

specific recombinogenic sequences, such as the imperfect

palindromic structure that is present in the CaMV 35S

promoter (Kohli et al. 1999). This sequence has the ability

to adopt a cruciform secondary structure, which may

Plant Mol Biol (2010) 73:363–378 371

123

stimulate recombination events. Notably, many other pro-

moters contain palindromic sequences of variable length

within 100 bp of the transcription start site, often because

they bind dimeric transcription factors, so it is possible that

promoters may be disproportionately involved in transgene

rearrangement events, a hypothesis that will need to be

examined through the detailed analysis of transgenic pop-

ulations carrying multiple copies of other promoters.

Evidence from many transformation experiments indi-

cates that there is no simple correlation between transgene

copy number and expression level, with the exception of

certain carefully controlled experiments using boundary

elements. In some cases, higher copy numbers have sup-

pressed overall expression levels (e.g., Cannell et al. 1999;

Spencer et al. 1992) whereas in others higher copy numbers

have enhanced expression (Stoger et al. 1998; Gahakwa

et al. 2000). Where suppression effects have occurred, it has

been suggested that ‘‘runaway expression’’ resulting in the

generation of aberrant RNAs lacking polyadenylate tails

has triggered potent silencing through the post-transcrip-

tional silencing pathway (Schubert et al. 2004). From the

prospect of promoter usage, this means in a practical sense

that multiple copies of the same promoter can appear in a

transgenic locus because of the presence of multiple copies

of the same transgene, so why should it matter if the pro-

moter is also represented with different transgenes? A

potential answer to this question comes from experiments

involving the use of the CaMV 35S promoter for large-scale

screening, and reveals a fourth mechanism that may trigger

transcriptional silencing. The CaMV 35S promoter and

enhancer have been used as a random insertional mutagen

to hyperactivate adjacent genes and generate gain-of-func-

tion phenotypes. It has been noted that such screens using

T-DNA cassettes containing the enhancer elements from

the CaMV 35S promoter return a lower than expected fre-

quency of morphological mutants (Chalfun et al. 2003).

Detailed analysis revealed a correlation between the num-

ber of T-DNA insertion sites, the methylation status of the

enhancer sequence and enhancer activity. All plants con-

taining more than a single T-DNA insert were methylated

on the enhancer and its activity was reduced, with the

amount of methylation and the reduction of enhancer

activity correlating with the number of T-DNA copies,

particularly those with right border inverted repeats (Chal-

fun et al. 2003). Even so, methylation was still detected at a

lower frequency in plants without right border inverted

repeats suggesting other triggers were active in these lines.

Alternative solutions

Research into the structure and activity of plant promoters

has allowed the identification of specific cis-acting elements

with defined functions, so the next generation of promoters

is likely to comprise a mix and match of cis-acting elements

that confer upon transgenes exactly the properties required

by the experimenter. To a certain extent this has already

been achieved through the use of heterologous recombinant

promoters such as the inducible promoters based on insect

hormones (see above), but the key disadvantage of these

systems is that they require the corresponding trans-acting

factors to be imported, a problem that would not be

encountered when using recombined endogenous elements

to generate bespoke plant promoters. In some cases, the aim

is simply to increase the activity of an endogenous pro-

moter, which can be achieved in the case of the CaMV 35S

promoter by duplicating its enhancer elements, a strategy

that also works when these elements are imported into other

promoters. Chimeric promoters have also been engineered

using sequences from the CoYMV and CsVMV major

transcript promoters as activating sequences to augment the

CaMV 35S promoter, resulting in higher activity compared

to CaMV 35S in stably transformed tobacco plants and

higher activity compared to corn c-zein in corn endosperm

(Rance et al. 2002). In other cases, it is desirable to con-

strain promoter activity in some way, which has been

achieved with synthetic auxin response elements that make

heterologous promoters more auxin-sensitive. In such cases,

the TGTCTC element is placed either as tandem or inverted

repeats with appropriate spacing, making the promoters up

to 10 times more sensitive to exogenous auxins (Guilfoyle

1999). The ‘super c-zein promoter’ contains a duplication

(region -444 to -174) from the corn endosperm-specific

27 kDa c-zein promoter, making it more active in endo-

sperm tissue (Marzabal et al. 1998). This has been used to

increase the expression of the bacterial genes crtB and crtI

resulting in greater accumulation of b-carotene (Aluru et al.

2008). Many plant promoters can also be enhanced by

including an intron (e.g., Vain et al. 1996; Mitsuhara et al.

1996; Fiume et al. 2004; Chiera et al. 2007).

Can this strategy of modulating promoter structure be

used to reduce promoter duplication in MGT? Several

reports have described the creation of artificial promoters

specifically to avoid repetition. For example, Sawant et al.

(2001) have described a strategy in which the sequences of

the strongest plant promoters were compared, to facilitate

the design of ideal artificial consensus promoters. By

combining the ideal minimal cassette (TATA box, initiator

and Kozak sites) with the ideal upstream activator, they

were able to develop a 450-bp synthetic promoter that was

active in cotton leaves, potato tubers and cabbage stems,

and was stronger than the CaMV 35S promoter in trans-

genic tobacco plants. In theory, several such artificial pro-

moters could be developed and used in combination, or with

native promoters. Bhullar et al. (2003) have developed two

alternative approaches to avoid promoter homology, one in

372 Plant Mol Biol (2010) 73:363–378

123

which functional cis-acting elements are embedded in dif-

ferent synthetic DNA sequences, and another in which the

functional cis-acting elements of different promoters with

the same activity are swapped to create non-homologous

chimeras. A set of modified CaMV 35S promoters was

developed and tested in transgenic tobacco plants, showing

that the chimeric promoters were generally weaker than the

native CaMV 35S promoter but that promoters constructed

from cis-elements embedded in a synthetic DNA sequence

were generally comparable in strength.

Most promoters used in plant biotechnology are unidi-

rectional, but bidirectional promoters are becoming

increasingly useful for MGT experiments as they allow the

simultaneous expression of two gene products. In many

promoters, only the core promoter sequence needs to be

oriented with respect to the coding sequence of the gene

to ensure transcription in the correct direction. Often,

upstream promoter elements can be present in either ori-

entation, and enhancer elements enjoy even more freedom

in terms of position and orientation. If a minimal promoter

consisting of a TATA box and initiator is duplicated and

placed facing outwards, cis-acting elements between the

minimal promoters can be used to drive transcription in

two directions simultaneously. The duplicated minimal

CaMV 35S promoter has been used for bidirectional tran-

scription in several different species (Xie et al. 2001;

Zhang et al. 2008) and can also be combined with other

promoters or synthetic response elements to make bidi-

rectional chimeric regulatory complexes (Chaturvedi et al.

2006). Xie et al. (2001) fused the minimal CaMV 35S

promoter to the constitutive PClSV promoter on one side

and the JA-sensitive OPR1 promoter at the other, and fused

a minimal promoter derived from the senescence-specific

SAG12 gene to the 50 end of the CaMV 35S promoter.

Naturally bidirectional promoters have been reported in

plants, including those directing the oleosin and methionine

sulphoxide reductase genes in canola (Keddie et al. 1994;

Sadanandom et al. 1996), the Arabidopsis cab1 and cab2

genes (chlorophyll a/b-binding protein; Mitra et al. 2009)

and the rice Ocip1 and Ocpi2 genes (chymotrypsin prote-

ase inhibitor; Singh et al. 2009). Bioinformatic analysis of

Arabidopsis, rice and poplar genome sequences has iden-

tified several putatively paired genes whose promoters may

be useful in the future for bidirectional gene expression

constructs (Dhadi et al. 2009). The different promoter

strategies discussed above are summarized in Fig. 1.

Conclusions and outlook

MGT is here to stay and in the future it will be desirable,

even essential, to increase the number of simultaneously

introduced transgenes still further (Naqvi et al. 2009a).

Currently, the creation of transgenic plants with stable

transgene expression relies on a healthy dose of luck, with

the best performing plants being selected from large pop-

ulations, the remaining plants being discarded if they fail to

live up to expectations. As MGT experiments become more

ambitious, the recovery rates will start to suffer from the

law of diminishing returns. How many plants will need to

be screened to find a line that stably expresses 10–20 or

more different transgenes? This is a realistic outlook for the

forthcoming decade, and in order to achieve such ambitious

objectives it will be necessary to minimize the likelihood

of transgene silencing triggered by the repetitive use of the

same promoter sequence. This will be addressed by a

number of strategies, including promoter diversity (facili-

tated by the discovery and characterization of new pro-

moters with relevant expression profiles), the use of

synthetic and modified promoters to reduce the extent of

homology between transgenes, and the use of bidirectional

promoters to reduce the promoter to transgene ratio

(a)

(b)

(c)

(d)

(e)

(f)

Fig. 1 Strategies for promoter deployment in nuclear multigene

transformation. a Repetitious use of the same promoter is a

convenient approach that often works effectively, but some research-

ers have found that it increases the risk of homology-dependent

transcriptional gene silencing. The following alternative approaches

may therefore be chosen instead. b Promoter diversity involves the

use of different promoters for each transgene, although this depends

on availability and becomes more difficult with larger numbers of

transgenes. c Synthetic consensus promoters can be developed from

the conserved features of native promoter sequences, and can add to

the repertoire of available promoters (Sawant et al. 2001). d Chimeric

promoters comprising the functional segments of native promoters

can be prepared by domain swapping (e.g., Bisht et al. 2004) or by

importing specific functional motifs from one promoter into another

(e.g., introducing the CaMV 35S enhancer or actin1 intron into

heterologous promoters to increase activity, or introducing hormone

response elements into constitutive promoters to make them induc-

ible). e The functional cis-elements of natural promoters can be

embedded in a synthetic DNA sequence (e.g., Bhullar et al. 2003).

f Bidirectional promoters can halve the number of promoters required

for MGT. Strategies d, e and f can be combined in different ways to

generate functionalized unidirectional and bidirectional chimeric

promoters with potentially limitless diversity

Plant Mol Biol (2010) 73:363–378 373

123

(Fig. 1). Novel approaches, such as the use of episomal

constructs like plant artificial chromosomes will also play a

role in the future, although for the time being such systems

are limited to a few model species and are technically

demanding.

Acknowledgments Work in our laboratory is supported by the

Ministry of Science and Innovation, Spain (BFU2007-61413 and

BIO2007-30738-E), a European Research Council Advanced Grant

(BIOFORCE) to PC, and the CONSOLIDER Agrigenomics program

funded by MICINN, Spain.

References

Abbadi A, Domergue F, Bauer J, Napier JA, Welti R, Zahringer U,

Cirpus P, Heinz E (2004) Biosynthesis of very-long-chain

polyunsaturated fatty acids in transgenic oilseeds: constraints on

their accumulation. Plant Cell 16:2734–2748

Afolabi AS, Worland B, Snape JW, Vain P (2004) A large-scale study

of rice plants transformed with different T-DNAs provides new

insights into locus composition and T-DNA linkage configura-

tions. Theor Appl Genet 109:815–826

Albani D, Hammond-Kosack MCU, Smith C, Conlan S, Colot V,

Holdsworth M, Bevan MW (1997) The wheat transcriptional

activator SPA: a seed-specific bZIP protein that recognizes the

GCN4-like motif in the bifactorial endosperm box of prolamin

genes. Plant Cell 9:171–184

Aluru M, Xu Y, Guo R, Wang Z, Li S, White W, Wang K, Rodermel

S (2008) Generation of transgenic maize with enhanced provi-

tamin A content. J Exp Bot 59:3551–3562

An G, Costa MA, Ha SB (1990) Nopaline synthase promoter is

wound inducible and auxin inducible. Plant Cell 2:225–233

An YQ, McDowell JM, Huang S, McKinney EC, Chambliss S,

Meagher RB (1996) Strong, constitutive expression of the

Arabidopsis ACT2/ACT8 actin subclass in vegetative tissues.

Plant J 10:107–121

Atanassova R, Leterrier M, Gaillard C, Agasse A, Sagot E, Coutos-

Thevenot P, Delrot S (2003) Sugar-regulated expression of a

putative hexose transport gene in grape. Plant Physiol 131:326–

334

Ballas N, Wong LM, Ke M, Theologis A (1995) Two auxin-

responsive domains interact positively to induce expression of

the early indoleacetic acid-inducible gene PS-IAA4/5. Proc Natl

Acad Sci USA 92:3483–3487

Baszczynski CL, Barbour E, Miki B (1997) ALS3 promoter. US

Patent 5659026

Bhattacharyya S, Dey N, Maiti IB (2002) Analysis of cis-sequence of

subgenomic transcript promoter from the Figwort mosaic virusand comparison of promoter activity with the Cauliflower mosaicvirus promoters in monocot and dicot cells. Virus Res 90:47–62

Bhattacharyya S, Pattanaik S, Maiti IB (2003) Intron-mediated

enhancement of gene expression in transgenic plants using

chimeric constructs composed of the Peanut chlorotic streakvirus (PClSV) promoter-leader and the antisense orientation of

PClSV ORF VII (p7R). Planta 218:115–124

Bhomkar P, Upadhyay C, Saxena M, Muthusamy A, Prakash NS,

Sarin NB (2008) Salt stress alleviation in transgenic Vignamungo L. Hepper (blackgram) by overexpression of the glyox-

alase I gene using a novel Cestrum yellow leaf curling virus(CmYLCV) promoter. Mol Breed 22:169–181

Bhullar S, Chakravarthy S, Advani S, Datta S, Pental D, Burma PK

(2003) Strategies for development of functionally equivalent

promoters with minimum sequence homology for transgene

expression in plants: cis-elements in a novel DNA context versus

domain swapping. Plant Physiol 132:988–998

Bisht NC, Jagannath A, Gupta V, Kumar Burma P, Pental D (2004) A

two gene–two promoter system for enhanced expression of a

restorer gene (barstar) and development of improved fertility

restorer lines for hybrid seed production in crop plants. Mol

Breed 14:129–144

Bisht NC, Jagannath A, Burma PK, Pradhan AK, Pental D (2007)

Retransformation of a male sterile barnase line with the barstar

gene as an efficient alternative method to identify male sterile-

restorer combinations for heterosis breeding. Plant Cell Rep

26:727–733

Braithwaite KS, Geijskes RJ, Smith GR (2004) A variable region of

the sugarcane bacilliform virus (SCBV) genome can be used to

generate promoters for transgene expression in sugarcane. Plant

Cell Rep 23:319–326

Callis J, Raasch JA, Vierstras RD (1990) Ubiquitin extension proteins

of Arabidopsis thaliana. Structure, localization, and expression

of their promoters in transgenic tobacco. J Biol Chem 265:

12466–12493

Cannell ME, Doherty A, Lazzeri PA, Barcelo P (1999) A population

of wheat and tritordeum transformants showing a high degree of

marker gene stability and heritability. Theor Appl Genet 99:

772–784

Cao X, Jacobsen SE (2002) Locus-specific control of asymmetric and

CpNpG methylation by the DRM and CMT3 methyltransferase

genes. Proc Natl Acad Sci USA 99:16491–16498

Cazzonelli CI, McCallum EJ, Lee R, Botella JR (2005) Character-

ization of a strong, constitutive mung bean (Vigna radiata L.)

promoter with a complex mode of regulation in planta.

Transgenic Res 14:941–967

Chalfun A, Mes JJ, Mlynarova L, Aarts MGM, Angenent GC (2003)

Low frequency of T-DNA based activation tagging in Arabid-

opsis is correlated with methylation of CaMV 35S enhancer

sequences. FEBS Lett 555:459–463

Chaturvedi CP, Sawant SV, Kiran K, Mehrotra R, Lodhi N, Ansari

SA, Tuli R (2006) Analysis of polarity in the expression from a

multifactorial bidirectional promoter designed for high-level

expression of transgenes in plants. J Biotechnol 123:1–12

Chiera JM, Bouchard RA, Dorsey SL, Park E, Buenrostro-Nava MT,

Ling PP, Finer JJ (2007) Isolation of two highly active soybean

(Glycine max (L.) Merr.) promoters and their characterization

using a new automated image collection and analysis system.

Plant Cell Rep 26:1501–1509

Christensen AH, Quail PH (1996) Ubiquitin promoter-based vectors

for high-level expression of selectable and/or screenable marker

genes in monocotyledonous plants. Transgenic Res 5:213–218

Christensen AH, Sharrock RA, Quail PH (1992) Maize polyubiquitin

genes: structure, thermal perturbation of expression and tran-

script splicing, and promoter activity following transfer to

protoplasts by electroporation. Plant Mol Biol 18:675–689

Colot V, Robert LS, Kavanagh TA, Bevan MW, Thompson RD

(1987) Localization of sequences in wheat endosperm protein

genes which confer tissue-specific expression in tobacco. EMBO

J 6:3559–3564

Cornejo MJ, Luth D, Blankenship KM, Anderson OD, Blechl AE

(1993) Activity of a maize ubiquitin promoter in transgenic rice.

Plant Mol Biol 23:567–581

De Block M, Debrouwer D (1991) Two T-DNAs co-transformed into

Brassica napus by a double Agrobacterium tumefaciens infec-

tion are mainly integrated at the same locus. Theor Appl Genet

82:257–263

De Buck S, Jacobs A, Van Montagu M, Depicker A (1999) The DNA

sequences of T-DNA junctions suggest that complex T-DNA

loci are formed by a recombination process resembling T-DNA

integration. Plant J 20:295–304

374 Plant Mol Biol (2010) 73:363–378

123

Dey N, Maiti IB (1999) Structure and promoter/leader deletion

analysis of mirabilis mosaic virus (MMV) full-length tran-

script promoter in transgenic plants. Plant Mol Biol 40:

771–782

Dhadi SR, Krom N, Ramakrishna W (2009) Genome-wide compar-

ative analysis of putative bidirectional promoters from rice,

Arabidopsis and Populus. Gene 429:65–73

Ellis JG, Tokuhisa JG, Llewellyn DJ, Bouchez D, Singh K, Dennis

ES, Peacock WJ (1993) Does the ocs-element occur as a

functional component of the promoters of plant genes? Plant J

4:433–443

Fang RX, Nagy F, Sivasubramaniam S, Chua NH (1989) Multiple cis-

regulatory elements for maximal expression of the Cauliflowermosaic virus 35S promoter in transgenic plants. Plant Cell

1:141–150

Farre G, Ramessar K, Twyman RM, Capell T, Christou P (2010) The

humanitarian impact of plant biotechnology: recent break-

throughs vs bottlenecks for adoption. Curr Opin Plant Biol (in

press; doi:10.1016/j.pbi.2009.11.002)

Fiume E, Christou P, Giani S, Breviario D (2004) Introns are key

regulatory elements of rice tubulin expression. Planta 218:

693–703

Fujisawa M, Takita E, Harada H, Sakurai N, Suzuki H, Ohyama K,

Shibata D, Misawa N (2009) Pathway engineering of Brassicanapus seeds using multiple key enzyme genes involved in

ketocarotenoid formation. J Exp Bot 60:1319–1332

Furtado A, Henry RJ (2005) The wheat Em promoter drives reporter

gene expression in embryo and aleurone tissue of transgenic

barley and rice. Plant Biotechnol J 3:421–434

Gahakwa D, Maqbool SB, Fu X, Sudhakar D, Christou P, Kohli A

(2000) Transgenic rice as a system to study the stability of

transgene expression: multiple heterologous transgenes show

similar behaviour in diverse genetic backgrounds. Theor Appl

Genet 101:388–399

Garbarino JE, Oosumi T, Belknap WR (1995) Isolation of a

polyubiquitin promoter and its expression in transgenic potato

plants. Plant Physiol 109:1371–1378

Godard KA, Byun-McKay A, Levasseur C, Plant A, Seguin A,

Bohlmann J (2007) Testing of a heterologous, wound- and

insect-inducible promoter for functional genomics studies in

conifer defense. Plant Cell Rep 26:2083–2090

Goddijn OJM, Lindsey K, van der Lee FM, Klap JC, Sijmons PC

(1993) Differential gene expression in nematode induced feeding

structures of transgenic plant harbouring promoter-gusA fusion

constructs. Plant J 4:863–873

Grierson C, Du JS, de Torres Zabala M, Beggs K, Smith C,

Holdsworth M, Bevan M (1994) Separate cis sequences and

trans factors direct metabolic and developmental regulation of a

potato tuber storage protein gene. Plant J 5:815–826

Guevara-Garcıa A, Mosqueda-Cano G, Arguello-Astorga G, Simpson

J, Herrera-Estrella L (1993) Tissue-specific and wound-inducible

pattern of expression of the mannopine synthase promoter is

determined by the interaction between positive and negative cis-

regulatory elements. Plant J 4:495–505

Guilfoyle TJ (1999) Auxin-regulated genes and promoters. In:

Hooykaas PJJ, Hall M, Libbenga KL (eds) Biochemistry

and molecular biology of plant hormones. Elsevier, London,

pp 423–459

He C, Lin Z, McElroy D, Wu R (2009a) Identification of a rice actin2gene regulatory region for high-level expression of transgenes in

monocots. Plant Biotechnol J 7:227–239

He XJ, Hsu YF, Pontes O, Zhu J, Lu J, Bressan RA, Pikaard C, Wang

CS, Zhu JK (2009b) NRPD4, a protein related to the RPB4

subunit of RNA polymerase II, is a component of RNA

polymerases IV and V and is required for RNA-directed DNA

methylation. Genes Dev 23:318–330

Hermann SR, Harding RM, Dale JL (2001) The banana actin 1

promoter drives near-constitutive transgene expression in veg-

etative tissues of banana (Musa spp.). Plant Cell Rep 20:525–530

Herr AJ, Jensen MB, Dalmay T, Baulcombe DC (2005) RNA

polymerase IV directs silencing of endogenous DNA. Science

308:118–120

Jeon J-S, Chung Y-Y, Lee S, Yi G-H, Oh B-G, An G (1999) Isolation

and characterization of an anther-specific gene, RA8, from rice

(Oryza sativa, L). Plant Mol Biol 39:35–44

Jones JDG, Gilbert DE, Grady KL, Jorgensen RA (1987) T-DNA

structure and gene expression in petunia plants transformed by

Agrobacterium tumefaciens C58 derivates. Mol Gen Genet

207:478–485

Jorgensen RA, Snyder C, Jones JDG (1987) T-DNA is organized

predominantly in inverted repeat structures in plants transformed

with Agrobacterium tumefaciens C58 derivatives. Mol Gen

Genet 207:471–477

Jose-Estanyol M, Perez P, Puigdomenech P (2005) Expression of the

promoter of HyPRP, an embryo-specific gene from Zea mays in

maize and tobacco transgenic plants. Gene 356:146–152

Kanno T, Mette MF, Kreil DP, Aufsatz W, Matzke M, Matzke AJ

(2004) Involvement of putative SNF2 chromatin remodeling

protein DRD1 in RNA-directed DNA methylation. Curr Biol

14:801–805

Kanno T, Huettel B, Mette MF, Aufsatz W, Jaligot E, Daxinger L,

Kreil DP, Matzke M, Matzke AJ (2005) Atypical RNA

polymerase subunits required for RNA-directed DNA methyla-

tion. Nature Genet 37:761–765

Karunanandaa B, Qi Q, Hao M, Baszis SR, Jensen PK, Wong YH,

Jiang J, Venkatramesh M, Gruys KJ, Moshiri F, Post-Beittenm-

iller D, Weiss JD, Valentin HE (2005) Metabolically engineered

oilseed crops with enhanced seed tocopherol. Metab Eng 7:

384–400

Kawalleck P, Somssich IE, Feldbrugge M, Hahlbrock K, Weisshaar B

(1993) Polyubiquitin gene expression and structural properties of

the ubi 2–4 gene in Petroselinum crispum. Plant Mol Biol

21:673–684

Kay R, Chan A, Daly M, McPherson J (1987) Duplication of CaMV

35S promoter sequences creates a strong enhancer for plant

genes. Science 236:1299–1302

Keddie JS, Tsiantis M, Piffanelli P, Cella R, Hatzopoulos P, Murphy

DJ (1994) A seed-specific Brassica napus oleosin promoter

interacts with a G-box-specific protein and may be bi-directional.

Plant Mol Biol 24:327–340

Keil M, Sanchez-Serrano JJ, Willmitzer L (1989) Both wound-

inducible and tuber-specific expression are mediated by the

promoter of a single member of the potato proteinase inhibitor II

gene family. EMBO J 5:1323–1330

Keinonen-Mettala K, Pappinen A, Von Weissenberg K (1998)

Comparisons of the efficiency of some promoters in silver birch

(Betula pendula). Plant Cell Rep 17:356–361

Kloti A, He X, Potrykus I, Hohn T, Futterer J (2002) Tissue-specific

silencing of a transgene in rice. Proc Natl Acad Sci USA

99:10881–10886

Kohli A, Griffiths S, Palacios N, Twyman RM, Vain P, Laurie DA,

Christou P (1999) Molecular characterization of transforming

plasmid rearrangements in transgenic rice reveals a recombina-

tion hotspot in the CaMV 35S promoter and confirms the

predominance of microhomology-mediated recombination. Plant

J 17:591–601

Kohli A, Twyman RM, Abranches A, Wegel E, Shaw P, Christou P,

Stoger E (2003) Transgene integration, organization and inter-

action in plants. Plant Mol Biol 52:247–258

Koltunow AM, Truettner J, Cox KH, Walroth M, Goldberg RB

(1990) Different temporal and spatial gene expression patterns

occur during anther development. Plant Cell 2:1201–1224

Plant Mol Biol (2010) 73:363–378 375

123

Kooter J, Matzke MA, Meyer P (1999) Listening to the silent genes:

transgene silencing, gene regulation and pathogen control.

Trends Plant Sci 4:340–346

Kumar S, Fladung M (2000) Transgene repeats in aspen: molecular

characterisation suggests simultaneous integration of indepen-

dent T-DNAs into receptive hotspots in the host genome. Mol

Gen Genet 264:20–28

Kumar S, Fladung M (2002) Transgene integration in aspen:

structures of integration sites and mechanism of T-DNA

integration. Plant J 31:543–551

Lam E, Benfey PN, Gilmartin PM, Fang RX, Chua NH (1989) Site-

specific mutations alter in vitro factor and change promoter

expression pattern in transgenic plants. Proc Natl Acad Sci USA

86:7890–7894

Leisy DJ, Hnilo J, Zhao Y, Okita TW (1989) Expression of a rice

glutelin promoter in transgenic tobacco. Plant Mol Biol 14:

41–50

Lu CA, Lim EK, Yu SM (1998) Sugar response sequence in the

promoter of a rice alpha-amylase gene serves as a transcriptional

enhancer. J Biol Chem 273:10120–10131

Luo K, Zhang G, Deng W, Luo F, Qiu K, Pei Y (2008) Functional

characterization of a cotton late embryogenesis-abundant D113gene promoter in transgenic tobacco. Plant Cell Rep 27:707–717

Ma JK, Drake PM, Christou P (2003) The production of recombinant

pharmaceutical proteins in plants. Nat Rev Genet 4:794–805

Maiti IB, Shepherd RJ (1998) Isolation and expression analysis of

Peanut chlorotic streak caulimovirus (PClSV) full-length tran-

script (FLt) promoter in transgenic plants. Biochem Biophys Res

Commun 244:440–444

Mandel T, Fleming AJ, Krahenbuhl R, Kuhlemeier C (1995)

Definition of constitutive gene expression in plants: the trans-

lation initiation factor 4A gene as a model. Plant Mol Biol

29:995–1004

Mariani C, De Beuckeleer M, Truettner J, Leemans J, Goldberg RB

(1990) Induction of male sterility in plants by a chimeric

ribonuclease gene. Nature 347:737–741

Mariani C, Gossele V, De Beuckeleer M, De Block M, Goldberg RB,

De Greef W, Leemans J (1992) A chimaeric ribonuclease-

inhibitor gene restores fertility to male sterile plants. Nature

357:384–387

Marris C, Gallois P, Copley J, Kreis M (1988) The 50 flanking region

of a barley B hordein gene controls tissue and developmental

specific CAT expression in tobacco plants. Plant Mol Biol

10:359–366

Martinez A, Sparks C, Hart CA, Thompson J, Jepson I (1999)

Ecdysone agonist inducible transcription in transgenic tobacco

plants. Plant J 19:97–106

Marzabal P, Busk PK, Ludevid MD, Torrent M (1998) The bifactorial

endosperm box of gamma-zein gene: characterisation and

function of the Pb3 and GZM cis-acting elements. Plant J

16:41–52

McElroy D, Zhang W, Cao J, Wu R (1990) Isolation of an efficient

actin promoter for use in rice transformation. Plant Cell 2:

163–171

McElroy D, Blowers AD, Jenes B, Wu R (1991) Construction of

expression vectors based on the rice actin 1 (Act1) 50 region for

use in monocot transformation. Mol Gen Genet 231:150–160

Medberry SL, Lockhart BEL, Olszewski NE (1992) The Commelinayellow mottle virus promoter is a strong promoter in vascular and

reproductive tissues. Plant Cell 4:185–192

Mehlo L, Mazithulela G, Twyman RM, Boulton MI, Davies JW,

Christou P (2000) Structural analysis of transgene rearrange-

ments and effects on expression in transgenic maize plants

generated by particle bombardment. Maydica 45:277–287

Mette MF, van der Winden J, Matzke MA, Matzke AJM (1999)

Production of aberrant promoter transcripts contributes to

methylation and silencing of unlinked homologous promoters

in trans. EMBO J 18:241–248

Mette MF, Aufsatz W, van der Winden J, Matzke MA, Matzke

AJM (2000) Transcriptional silencing and promoter methyla-

tion triggered by double-stranded RNA. EMBO J 19:5194–

5201

Mitra A, Han J, Zhang ZJ, Mitra A (2009) The intergenic region of

Arabidopsis thaliana cab1 and cab2 divergent genes functions as

a bidirectional promoter. Planta 229:1015–1022

Mitsuhara I, Ugaki M, Hirochika H, Ohshima M, Murakami T, Gotoh

Y, Katayose Y, Nakamura S, Honkura R, Nishimiya S, Ueno K,

Mochizuki A, Tanimoto H, Tsugawa H, Otsuki Y, Ohashi Y

(1996) Efficient promoter cassettes for enhanced expression of

foreign genes in dicotyledonous and monocotyledonous plants.

Plant Cell Physiol 37:49–59

Mourrain P, van Blokland R, Kooter JM, Vaucheret H (2007) A

single transgene locus triggers both transcriptional and post-

transcriptional silencing through double-stranded RNA produc-

tion. Planta 225:365–379

Naqvi S, Farre G, Sanahuja G, Capell T, Zhu C, Christou P (2009a)

When more is better: multigene engineering in plants. Trends

Plant Sci 15:48–56

Naqvi S, Zhu C, Farre G, Ramessar K, Bassie L, Breitenbach J, Perez

Conesa D, Ros G, Sandmann G, Capell T, Christou P (2009b)

Transgenic multivitamin corn through biofortification of endo-

sperm with three vitamins representing three distinct metabolic

pathways. Proc Natl Acad Sci USA 106:7762–7767

Ni M, Cui D, Gelvin SB (1996) Sequence-specific interactions of

wound-inducible nuclear factors with mannopine synthase 20

promoter wound-responsive elements. Plant Mol Biol 30:

77–96

Norris SR, Meyer SE, Callis J (1993) The intron of Arabidopsisthaliana polyubiquitin genes is conserved in location and is a

quantitative determinant of chimeric gene expression. Plant Mol

Biol 21:895–906

Odell JT, Nagy F, Chua NH (1985) Identification of DNA sequences

required for activity of the Cauliflower mosaic virus 35S

promoter. Nature 313:810–812

Ohtsuki S, Levine M, Cai HN (1998) Different core promoters

possess distinct regulatory activities in the Drosophila embryo.

Genes Dev 12:547–556

Onate L, Vicente-Carbajosa J, Lara P, Dıaz I, Carbonero P (1999)

Barley BLZ2, a seed-specific bZIP protein that interacts with

BLZ1 in vivo and activates transcription from the GCN4-like

motif of B-hordein promoters in barley endosperm. J Biol Chem

274:9175–9182

Onodera Y, Suzuki A, Wu CY, Washida H, Takaiwa F (2001) A rice

functional transcriptional activator, RISBZ1, responsible for

endosperm-specific expression of storage protein genes through

GCN4 motif. J Biol Chem 276:14139–14152

Onodera Y, Haag JR, Ream T, Nunes PC, Pontes O, Pikaard CS

(2005) Plant nuclear RNA polymerase IV mediates siRNA and

DNA methylation-dependent heterochromatin formation. Cell

120:613–622

Opsahl-Sorteberg HG, Divon HH, Nielsen PS, Kalla R, Hammon-

Kosach M, Shimamoto K, Kohli A (2004) Identification of a 49-

bp fragment of the HvLTP2 promoter directing aleurone cell

specific expression. Gene 341:49–58

Padidam M, Gore M, Lu DL, Smirnova O (2003) Chemical-inducible,

ecdysone receptor-based gene expression system for plants.

Transgenic Res 12:101–109

Panikulangara TJ, Eggers-Schumacher G, Wunderlich M, Stransky H,

Schoffl F (2004) Galactinol synthase1. A novel heat shock factor

target gene responsible for heat-induced synthesis of raffinose

family oligosaccharides in Arabidopsis. Plant Physiol 136:3148–

3158

376 Plant Mol Biol (2010) 73:363–378

123

Paul W, Hodge R, Smartt S, Draper J, Scott R (1992) The isolation

and characterization of the tapetum specific Arabidopsis thalianaA9 gene. Plant Mol Biol 19:611–622

Pontier D, Yahubyan G, Vega D, Bulski A, Saez-Vasquez J, Hakimi

MA, Lerbs-Mache S, Colot V, Lagrange T (2005) Reinforce-

ment of silencing at transposons and highly repeated sequences

requires the concerted action of two distinct RNA polymerases

IV in Arabidopsis. Genes Dev 19:2030–2040

Potenza C, Aleman L, Sengupta-Gopalan C (2004) Targeting

transgene expression in research, agricultural, and environmental

applications: promoters used in plant transformation. In Vitro

Cell Dev Biol Plant 40:1–22

Qu LQ, Takaiwa F (2004) Evaluation of tissue specificity and

expression strength of rice seed component gene promoters in

transgenic rice. Plant Biotechnol J 2:113–125

Raclaru M, Gruber J, Kumar R, Sadre R, Luhs W, Zarhloul MK,

Friedt W, Frentzen M, Weier D (2006) Increase of the

tocochromanol content in transgenic Brassica napus seeds by

overexpression of key enzymes involved in prenylquinone

biosynthesis. Mol Breed 18:93–107

Raho G, Lupotto E, Hartings H, Della Torre AP, Perrotta C,

Marmiroli N (1996) Tissue-specific expression and environmen-

tal regulation of the barley Hvhspi7 gene promoter in transgenic

tobacco plants. J Exp Bot 47:1587–1594

Rai M, Datta K, Parkhi V, Tan J, Oliva N, Chawla HS, Datta SK

(2007) Variable T-DNA linkage configuration affects inheritance

of carotenogenic transgenes and carotenoid accumulation in

transgenic indica rice. Plant Cell Rep 26:1221–1231

Rance I, Norre F, Gruber V, Theisen M (2002) Combination of viral

promoter sequences to generate highly active promoters for

heterologous therapeutic protein over-expression in plants. Plant

Sci 162:833–842

Rushton PJ, Reinstadler A, Lipka V, Lippok B, Somssich IE (2002)

Synthetic plant promoters containing defined regulatory ele-

ments provide novel insights into pathogen- and wound-induced

signaling. Plant Cell 14:749–762

Sadanandom A, Piffanelli P, Knott T, Robinson C, Sharpe A, Lydiate

D, Murphy D, Fairbairn DJ (1996) Identification of a peptide

methionine sulphoxide reductase gene in an oleosin promoter

from Brassica napus. Plant J 10:235–242

Saidi Y, Finka A, Chakhporanian M, Zryd JP, Schaefer DG,

Goloubinoff P (2005) Controlled expression of recombinant

proteins in Physcomitrella patens by a conditional heat-shock

promoter: a tool for plant research and biotechnology. Plant Mol

Biol 59:697–711

Sasaki K, Yuichi O, Hiraga S, Gotoh Y, Seo S, Mitsuhara I, Ito H,

Matsui H, Ohashi Y (2007) Characterization of two rice

peroxidase promoters that respond to blast fungus-infection.

Mol Genet Genomics 278:709–722

Sawant S, Singh PK, Madanala R, Tuli R (2001) Designing of an

artificial expression cassette for the high-level expression of

transgenes in plants. Theor Appl Genet 102:635–644

Schenk PM, Remans T, Sagi L, Elliott AR, Dietzgen RG, Swennen R,

Ebert PR, Grof CPL, Manners JM (2001) Promoters for

pregenomic RNA of banana streak badnavirus are active for

transgene expression in monocot and dicot plants. Plant Mol Biol

47:399–412

Schernthaner JP, Matzke MA, Matzke AJM (1988) Endosperm-

specific activity of a zein gene promoter in transgenic tobacco

plants. EMBO J 7:1249–1255

Schmidt RJ, Ketudat M, Aukerman MJ, Hoschek G (1992) Opaque-2

is a transcriptional activator that recognizes a specific target site

in 22-kD zein genes. Plant Cell 4:689–700

Schubert D, Lechtenberg B, Forsbach A, Gils M, Bahadur S, Schmidt

R (2004) Silencing in Arabidopsis T-DNA transformants: the

predominant role of a gene-specific RNA sensing mechanism

versus position effects. Plant Cell 16:2561–2572

Schunmann PHD, Llewellyn DJ, Surin B, Boevink P, Feyter RCD,

Waterhouse PM (2003a) A suite of novel promoters and

terminators for plant biotechnology. Functional Plant Biol

30:443–452

Schunmann PHD, Surin B, Waterhouse PM (2003b) A suite of novel

promoters and terminators for plant biotechnology. II. The

pPLEX series for use in monocots. Functional Plant Biol

30:453–460

Singh A, Sahi C, Grover A (2009) Chymotrypsin protease inhibitor

gene family in rice: genomic organization and evidence for the

presence of a bidirectional promoter shared between two

chymotrypsin protease inhibitor genes. Gene 428:9–19

Skriver K, Olsen FL, Rogers JC, Mundy J (1991) Cis-acting DNA

elements responsive to gibberellin and its antagonist abscisic

acid. Proc Natl Acad Sci USA 88:7266–7270

Slater A, Scott N, Fowler M (2003) Plant Molecular Biology. Oxford

University Press, Oxford

Spencer TM, O’Brien JV, Start WG, Adams TR, Gordon-Kamm WJ,

Lemaux PG (1992) Segregation of transgenes in maize. Plant

Mol Biol 18:201–210

Stavolone L, Ragozzino A, Hohn T (2003) Characterization of

Cestrum yellow leaf curling virus: a new member of the

caulimoviridae family. J Gen Virol 84:3459–3464

Stoger E, Williams S, Keen D, Christou P (1998) Molecular

characteristics of transgenic wheat and the effect on transgene

expression. Transgenic Res 7:463–471

Sun TP, Gubler F (2004) Molecular mechanism of gibberellin

signaling in plants. Annu Rev Plant Biol 55:197–223

Sun AQ, Yi SY, Yang JY, Zhao CM, Liu J (2006) Identification and

characterization of a heat-inducible ftsH gene from tomato

(Lycopersicon esculentum Mill.). Plant Sci 170:551–562

Takaiwa F, Oono K, Kato A (1991) Analysis of the 50 flanking region

responsible for the endosperm-specific expression of a rice

glutelin chimeric gene in transgenic tobacco. Plant Mol Biol

16:49–58

Takaiwa F, Yamanouchi U, Yoshihara T, Washida H, Tanabe F, Kato

A, Yamada K (1996) Characterization of common cis-regulatory

elements responsible for the endosperm-specific expression of

members of the rice glutelin multigene family. Plant Mol Biol

30:1207–1221

Tiana L, Wub K, Hannama C, Latoszek-Greena M, Sibbalda S, Huc

M, Browna DCW, Mikic B (2005) Analysis and use of the

tobacco eIF4A–10 promoter elements for transgene expression. J

Plant Physiol 162:1355–1366

Torbert KA, Gopalraj M, Medberry SL, Olszewski NE, Somers DA

(1998) Expression of the Commelina yellow mottle viruspromoter in transgenic oat. Plant Cell Rep 17:284–287

Twyman RM (2003) Growth and development: control of gene

expression, regulation of transcription. In: Thomas B, Murphy

DJ, Murray B (eds) Encyclopedia of Applied Plant Sciences.

Elsevier Science, London, pp 558–567

Unger E, Cigan AM, Trimnell M, Xu RJ, Kendall T, Roth B,

Albertsen M (2002) A chimeric ecdysone receptor facilitates

methoxyfenozide-dependent restoration of male fertility in ms45

maize. Transgenic Res 11:455–465

Urwin PE, Moller SG, Lilley CJ, McPherson MJ, Atkinson HJ (1997)

Continual green fluorescent protein monitoring of Cauliflowermosaic virus 35S promoter activity in nematode-induced feeding

cells in Arabidopsis thaliana. Mol Plant Microbe Interact

10:394–400

Vain P, Finer KR, Engler DE, Pratt RC, Finer JJ (1996) Intron-

mediated enhancement of gene expression in maize (Zea maysL.) and bluegrass (Poa pratensis L.). Plant Cell Rep 15:489–494

Plant Mol Biol (2010) 73:363–378 377

123

Valentin HE, Mitsky TA (2002) TYRA genes and uses thereof. Patent

Application WO 02/089561 A1

Veena, Reddy VS, Sopory SK (1999) Glyoxalase I from Brassicajuncea: molecular cloning, regulation and its over-expression

confer tolerance in transgenic tobacco under stress. Plant J

17:385–395

Verdaguer B, de Kochko A, Beachy RN, Fauquet C (1996) Isolation

and expression in transgenic tobacco and rice plants, of the

Cassava vein mosaic virus (CsVMV) promoter. Plant Mol Biol

31:1129–1139

Vicente-Carbajosa J, Onate L, Lara P, Diaz I, Carbonero P (1998)

Barley BLZ1: a bZIP transcriptional activator that interacts with

endosperm-specific gene promoters. Plant J 13:629–640

Wierzbicki AT, Haag JR, Pikaard CS (2008) Noncoding transcription

by RNA polymerase Pol IVb/Pol V mediates transcriptional

silencing of overlapping and adjacent genes. Cell 135:635–648

Wobus U, Borisjuk L, Panitz R, Manteuffel R, Baumlein H, Wohlfart

T, Heim U, Weber H, Misera S, Weschke W (1995) Control of

seed storage protein gene expression: new aspects on an old

story. J Plant Physiol 145:592–599

Woodger F, Millar A, Murray F, Jacobsen J, Gubler F (2003) The role

of GAMYB transcription factors in GA-regulated gene expres-

sion. J Plant Growth Regul 22:176–184

Wu CY, Suzuki A, Washida H, Takaiwa F (1998) The GCN4 motif in

a rice glutelin gene is essential for endosperm-specific gene

expression and is activated by Opaque-2 in transgenic rice

plants. Plant J 14:673–683

Wu H, Michler CH, LaRussa L, Davis JM (1999) The pine Pschi4promoter directs wound-induced transcription. Plant Sci

142:199–207

Xie M, He Y, Gan S (2001) Bidirectionalization of polar promoters in

plants. Nature Biotechnol 19:677–679

Xie Z, Johansen L, Gustafson A, Kasschau K, Lellis A, Zilberman D,

Jacobsen S, Carrington J (2004) Genetic and functional diver-

sification of small RNA pathways in plants. PLoS Biol 2:E104

Xu D, McElroy D, Thornburg RW, Wu R (1993) Systemic induction

of a potato pin2 promoter by wounding, methyl jasmonate, and

abscisic acid in transgenic rice plants. Plant Mol Biol 22:

573–588

Yang G, Nakamura H, Ichikawa H, Kitano H, Komatsu S (2006)

OsBLE3, a brassinolide-enhanced gene, is involved in the growth

of rice. Phytochemistry 67:1442–1454

Yevtushenko DP, Sidorov VA, Romero R, Kay WW, Misra S (2004)

Wound-inducible promoter from poplar is responsive to fungal

infection in transgenic potato. Plant Sci 167:715–724

Zhang W, McElroy D, Wu R (1991) Analysis of rice Act1 50 region

activity in transgenic rice plants. Plant Cell 3:1155–1165

Zhang C, Gai Y, Wang W, Zhu Y, Chen X, Jiang X (2008)

Construction and analysis of a plant transformation binary vector

pBDGG harboring a bi-directional promoter fusing dual visible

reporter genes. J Genet Genomics 35:245–249

Zheng Z, Kawagoe Y, Xiao S, Li Z, Okita TW, Hau TL, Lin A, Murai

N (1993) 50 Distal and proximal cis-acting regulation elements

are required for developmental control of a rice seed storage

protein glutelin gene. Plant J 4:357–366

Zheng X, Deng W, Luo K, Duan H, Chen Y, McAvoy R, Song S, Pei

Y, Li Y (2007) The Cauliflower mosaic virus (CaMV) 35S

promoter sequence alters the level and patterns of activity of

adjacent tissue- and organ-specific gene promoters. Plant Cell

Rep 26:1195–1203

Zhu QH, Ramm K, Eamens AL, Dennis ES, Upadhyaya NM (2006)

Transgene structures suggest that multiple mechanisms are

involved in T-DNA integration in plants. Plant Sci 171:308–322

Zhu C, Naqvi S, Gomez-Galera S, Pelacho AM, Capell T, Christou P

(2007) Transgenic strategies for the nutritional enhancement of

plants. Trends Plant Sci 12:548–555

Zhu C, Naqvi S, Breitenbach J, Sandmann G, Christou P, Capell T

(2008) Combinatorial genetic transformation generates a library

of metabolic phenotypes for the carotenoid pathway in maize.

Proc Natl Acad Sci USA 105:18232–18237

Zilberman D, Cao X, Jacobsen SE (2003) ARGONAUTE4 control of

locus-specific siRNA accumulation and DNA and histone

methylation. Science 299:716–719

Zourelidou M, de Torres-Zabala M, Smith C, Bevan MW (2002)

Storekeeper defines a new class of plant-specific DNA-binding

proteins and is a putative regulator of patatin expression. Plant J

30:489–497

378 Plant Mol Biol (2010) 73:363–378

123